content
stringlengths
1
15.9M
\section{Introduction} Ultracold atoms in optical lattices serve now as a routine tool to study various lattice models derived from other areas of physics, such as condensed matter or high energy physics. They often enrich the original models with additional features accessible due to extreme controllability and versatility of possible experimental realizations (for recent reviews see \cite{Lewenstein07,Bloch08,Lewenstein12}). This potential of cold atoms was recognized for the first time thanks to a seminal theoretical proposal of Jaksch {\it et al.} \cite{Jaksch98}. Soon followed the experimental demonstration of the superfluid-Mott insulator phase transition \cite{Greiner02}, and the intensive studies of the effects of disorder on cold-atom systems \cite{Damski03,Roth03}. After a series of attempts \cite{Lye05,Fort05,Schulte05,Clement05,Clement06,Schulte06,Lye07} the Anderson localization for non-interacting (expanding) atoms was unambiguously demonstrated in \cite{Billy08}, The experimental studies of the phase diagram of interacting bosonic atoms in a disordered potential revealed the long-debated gapless insulating Bose glass (BG) phase \cite{Fallani07}, stimulating at the same time the discussion on the ways to observe and detect the BG phase \cite{Roux08,Roscilde08,Zakrzewski09,Delande09,Roux13}. Various aspects of Anderson localization in cold atoms are reviewed in Ref. \cite{Aspect09,Modugno10}, while the importance of disorder studies in a broader context of quantum simulators is presented in \cite{LSPLew10}. Optical implementations of quenched disorder are unique in the sense that the disorder can be, in principle, controlled with high precision on demand. Various methods of creating disorder have been proposed. A speckle pattern collimated on the atomic sample \cite{Damski03} gives rise to a truly random intensity landscape, which follows the exponential distribution. Another proposed scheme \cite{Damski03,Roth03} applies several (at least two) laser fields with different frequencies. For an appropriate choice of frequencies the resulting potential is quasi-periodic, and for a finite sample hardly distinguishable from a truly random case (see \cite{Roati08}, and also \cite{Orso09,DErrico14,Tanzi13} for recent results). Yet another interesting way to create the disordered potential is to use the interactions between atoms. By pinning a secondary type of atoms in an optical potential we obtain the disorder with binary (or Bernoulli) distribution. Such proposal, originating from the work of Gavish and Castin \cite{Gavish05}, has been frequently discussed theoretically \cite{Massignan06,Mering2008pra,Krutitsky2008pra,Buonsante2009pra, Stasinska2012njp}, and only recently implemented experimentally \cite{Gadway2011prl} (note that in the early papers \cite{Ospelkaus2006prl} the impurities were mobile). This is the type of disorder we consider in this paper. Specifically, we study the bosonic atoms in the optical lattice potential, which interact with immobile randomly distributed atoms of a secondary (fermionic) species. Such a situation is routinely described using the Bose-Hubbard model with the random potential \begin{equation}\label{eq:Hamiltonian_1} H_{1}=-t\sum_{\langle i,j \rangle} b_i^{\dagger} b_j +\frac{U}{2} \sum_i b_i^{\dagger}b_i^{\dagger} b_i b_i-\sum_i \left(\mu-\gamma\omega_i\right) b_i^{\dagger}b_i, \end{equation} where $b_i, b_i^{\dagger}$ are the bosonic annihilation and creation operators, $t$ is the tunneling, the interaction constant is denoted by $U$, $\mu$ is the chemical potential and $\langle i,j \rangle$ denotes the summation over the nearest neighbors. The parameter $\gamma$ characterizes the strength of the disorder and $\omega_i$ is a random variable with binary distribution (taking value $1$ with probability $p$ when the heavy background fermion is present at the $i$th site, and $0$ with probability $(1-p)$, when no impurity is present). This Hamiltonian was already studied in several works \cite{Mering2008pra,Krutitsky2008pra,Buonsante2009pra}, where the emergence of the Mott insulating (MI) phase with non-integer filling related to the impurity density was demonstrated, and the MI phase was shown to survive for arbitrarily strong disorder unlike in the continuous disorder case. The Bose-Hubbard model with different forms of diagonal disorder has been reviewed in \cite{Pollet13} while random on-site interactions have been considered in \cite{Gimperlein05} It has been shown recently that the simple Bose-Hubbard description for fermion-boson mixture may not be adequate for stronger interspecies interactions. The shift of the observed transition between the superfluid (SF) and the MI phases, observed experimentally in Refs. \cite{Ospelkaus2006prl,Gunter06,Best09} (for the analogous effects with Bose-Bose mixtures and tightly trapped bosons, see \cite{Thalhammer2008,Will2010}), could not be explained with this simple description. It has soon been realized that density-dependent tunneling terms as well as contributions coming from higher Bloch bands are necessary to describe the systems in question \cite{Luhmann08,Mering2011pra,Jurgensen12,Bissbort2012pra}, whenever the interspecies interaction becomes strong enough. As a consequence, for bosons interacting with immobile fermionic impurities the disorder affects also the tunneling. While similar corrections could be also taken into account for boson-boson interactions, we do not include them to simplify the picture, assuming boson-boson interactions to be sufficiently weak (see a recent review \cite{review} for discussion of different possible contributions). In this simplified picture we add to the Hamiltonian (\ref{eq:Hamiltonian_1}) the term $T(\omega_i+\omega_j)] b_i^{\dagger} b_j$ yielding the tunneling dependent on the presence of the heavy fermion. The density induced tunneling coefficient $T$, proportional to boson-fermion interaction strength, seems at first independent of the standard tunneling $t$. However, this is not the case in the optical lattice potential, where both $t$ and $T$ depend on the potential depth (when proper Wannier functions are used to evaluate them), and are (see e.g. \cite{review}) approximately proportional to each other for standard depths of optical lattices. Thus, we may assume $T=\alpha t$ obtaining the Hamiltonian: \begin{eqnarray}\label{eq:Hamiltonian_2} H_{2}&=&-t\sum_{\langle i,j \rangle} [1+\alpha(\omega_i+\omega_j)] b_i^{\dagger} b_j\\ &&+\frac{U}{2} \sum_i b_i^{\dagger}b_i^{\dagger} b_i b_i-\sum_i \left(\mu-\gamma\omega_i\right) b_i^{\dagger}b_i.\nonumber \end{eqnarray} Here, $\alpha$ and $ \gamma$ depend on the interaction between the two species and, typically, $|\gamma| > |\alpha|$ \cite{review}. They are of the same sign, and below we only discuss the case of positive $\alpha$ and $\gamma$. This corresponds to the repulsive boson-fermion interaction. Note that in this model a single local random variable related to the presence of the impurity, enters in the potential and the tunneling terms of the Hamiltonian of the system. For the sake of comparison we shall also study a different model in which the disorder in the tunneling is given by independent random variables $\Omega_i\neq \omega_i$ (see also \cite{Grzybowski2006pss,Kruger2009prb} for models with simultaneous potential and hopping disorder): \begin{eqnarray}\label{eq:Hamiltonian_3} H_{3}&=&-t\sum_{<i,j>} [1+\alpha(\Omega_i+\Omega_j)] b_i^{\dagger} b_j\\ && +\frac{U}{2} \sum_i b_i^{\dagger}b_i^{\dagger} b_i b_i-\sum_i \left(\mu-\gamma\omega_i\right) b_i^{\dagger}b_i.\nonumber \end{eqnarray} The main goal of the present paper is to compare the phase diagrams of Hamiltonians $H_1, H_2$ and $H_3$. To this end we develop a method being a combination of a site-dependent decoupling mean-field method with a ``Hartree-Fock-like'' procedure \cite{Stasinska2014}. Independently we use also the Gutzwiller-ansatz approach. Note that the type of disorder that we study does not fulfill the assumptions of the "theorem of inclusions", valid for continuously distributed bounded disorder \cite{Pollet09}. We perform the analysis mostly in two dimensions (2D) as the mean-field approaches cannot be regarded as accurate in 1D. Our main results are: i) for the Hamiltonian $H_1$ there is a direct MI-SF transition at the tips of the Mott lobes in the thermodynamic limit; ii) for the Hamiltonian $H_2$ the Mott lobes are smaller, and the direct MI-SF transition at their tips disappears, although the region of BG is very narrow; iii) finally, for $H_3$ there is no direct MI-SF transition and the region of BG is wider in comparison to $H_2$. One of the main aspects of this paper is also the use of the mean-field theory combined with the simple Hartree-Fock approach, quite different from what has been proposed so far \cite{Fisher89,Sheshadri1995prl,Buonsante2007pra,Pisarski2011pra,Niederle2013,Bissbort2009epl,Bissbort2010pra}, and quite efficient in determining the phase boundary between the BG and the SF phase. The paper is structured as follows. In section \ref{sec:survey} we briefly recall the mean-field approaches applied in the studies of the disordered Bose-Hubbard model. In section \ref{sec:method} we introduce a method being a combination of a local mean-field approach and a Hartree-Fock-like method. The standard Gutzwiller approach used later for comparison is presented in \ref{sec:Gutzwiller}. In section \ref{sec:results} we apply the theory developed in section \ref{sec:method} to compare the phase diagrams of the bosonic atoms interacting with immobile impurities on a lattice described by the Hamiltonians $H_1$, $H_2$ and $H_3$ and confront these results with the Gutzwiller approach. Finally, we conclude in section \ref{sec:summary} by summarizing the obtained results. \section{Brief survey of mean-field approaches for the disordered Bose-Hubbard model}\label{sec:survey} The phase diagram of the disordered Bose-Hubbard model has been studied by a number of methods including the quantum Monte Carlo \cite{Krauth1991,Scalettar1991,Batrouni1992,Kisker1997,Prokofev2004}, renormalization group \cite{Singh1992,Svistunov1996}, density-matrix renormalization group techniques \cite{Pai1996,Rapsch1999,Lee2001}, tensor networks-based algorithms, or various mean-field approaches \cite{Fisher89,Sheshadri1995prl,Damski03,Krutitsky2006}. In this work we propose an extension of the local mean-field method, thus let us first briefly review the mean-field approaches used earlier. The local mean-field method was introduced in Ref. \cite{Fisher89}, and further developed in \cite{Buonsante2007pra}, where the boundary of the Mott lobe was linked to the stability of the zero solution of the self-consistency equations, which is then studied through linearization of those equations. Moreover the authors suggest that the presence of the BG surrounding the MI phase could be inferred from the spectral properties of the random matrix, which appear in the linearized problem. In \cite{Pisarski2011pra} the authors generalize the inhomogeneous site-dependent mean-field to clusters, which allows them to include the (short-range) spatial correlations. Sheshadri and co-workers \cite{Sheshadri1995prl} proposed to study the Bose-Hubbard model with disordered potential using the inhomogeneous generalization of the site-decoupling mean-field method, i.e., the local mean-field theory . There the hopping term is decoupled as $b_i^{\dagger}b_j\approx b_i^{\dagger}\av{b_j}+b_j\av{b_i^{\dagger}}-\av{b_i^{\dagger}}\av{b_j}$ yielding single site Hamiltonians coupled to neighbouring sites only through SF amplitudes $\psi_j=\av{b_j}$. The authors then diagonalize the local Hamiltonians in the occupation-number basis and determine self-consistently the values of $\psi_j$ minimizing the energy. The BG-SF transition is characterized through the percolation of sites with non-zero SF parameter. A description equivalent to the local mean-field theory may be achieved by minimizing the average energy over a variational manifold composed of products of single site wave vectors. Indeed, under such assumptions $\langle a_i a_j^\dagger\rangle = \langle a_i \rangle \langle a_j^\dagger\rangle$. This numerical ansatz is called the Gutzwiller ansatz. Yet another approach, the stochastic mean-field method, was proposed by Bissbort and Hofstetter in \cite{Bissbort2009epl} (see also \cite{Bissbort2010pra}). There, the starting point is also the decoupling of the tunneling term, however instead of choosing a different mean-field parameter for each site, the authors consider the probability distribution $P(\psi)$, reducing the description to effectively single-site problem. The probability distribution of $\psi$ is then found self-consistently to ensure the compatibility of $P(\psi)$ with the distribution on the neighbouring sites. Finally, in a recent work Niederle and Rieger \cite{Niederle2013} compare the results obtained with the local mean-field theory and the stochastic mean-field method with the quantum Monte-Carlo results. They conclude that the identification of different phases based on averaged quantities obtained through the mean-field approaches is misleading. Instead, the authors propose to distinguish the phases through the presence and percolation of the SF clusters finding an excellent agreement with the quantum Monte-Carlo studies. In what follows we will take the latter approach, i.e., study the percolation of the SF clusters; we propose, however, a different method to determine the distribution of the ``superfluid particles'' in the lattice. To that end we combine the mean field approach with Hartree-Fock-like method mixing different mean field modes. In this way, at a little numerical effort, we can go beyond the standard mean field approach. \section{Local mean-field approach combined with a Hartree-Fock-like method}\label{sec:method} The proposed method is a compromise between the local mean-field description which, based on the product-state description of the system, may not capture correctly the long-range correlations, and the resource-demanding numerical approaches. We also make an attempt to include the spatial correlations on top of the simple local mean-field description. Since the Hamiltonians $H_1$ and $H_2$ can be considered as special cases of the Hamiltonian $H_3$, in what follows we will concentrate on the latter. \subsection{Standard mean-field approach - local Hamiltonian} Like in the routine mean-field approach \cite{Fisher89} we begin by decoupling the Hamiltonian (\ref{eq:Hamiltonian_3}) using the standard approximation $b_i^{\dagger}b_j\approx b_i^{\dagger}\av{b_j}+b_j\av{b_i^{\dagger}}-\av{b_i^{\dagger}}\av{b_j}$ \cite{Sheshadri1993epl} and introducing a local mean-field parameter $\psi_i=\av{b_i}$. As a result we obtain: \begin{eqnarray} H&=&H_i+\bar{t} \sum_{\langle i,j \rangle} [1+\alpha(\Omega_i+\Omega_j)] \psi_i^{*} \psi_j,\label{eq:Hamiltonian}\\ H_i&=&-\bar{t}\sum_{\langle j\rangle_i}[1+\alpha(\Omega_i+\Omega_j)]\psi_j(b_i+b_i^{\dagger})\nonumber\\ &&+\frac{1}{2} b_i^{\dagger} b_i^{\dagger} b_i b_i-(\bar{\mu}-\bar{\gamma}\omega_i) b_i^{\dagger}b_i, \label{eq:Hamiltonian_i} \end{eqnarray} where we also express all the parameters in the units of $U$, i.e., $\bar{t}=t/U,\bar{\mu}=\mu/U,\bar{\gamma}=\gamma/U$. \subsection{Standard mean-field approach - energy minimization and self-consistency} The ground state, or more generally Gibbs energy of the local mean-field Hamiltonian is a highly nonlinear function of the local mean fields $ \psi_j$'s. The next step in the standard approach is to find a minimum of the energy under the constraint that $\av{b_j}= \psi_j$. This is in general a complicated task, but as long as we are interested in finding the boundaries of the MI phase, the analysis can be restricted to small values of $\psi_j$'s, where the energy is a quadratic form of $\psi_j$'s. The solutions of the self-consistency equations can be then obtained via a perturbative expansion up to first order in $\bar t$: \begin{equation}\label{eq:self_consistent} \psi_i:=\av{b_i}=\sum_{\langle j\rangle_i}\bar t \mathcal{R}_{ij} \psi_j, \end{equation} with the random matrix \begin{eqnarray} &&\mathcal{R}_{ij}=[1+\alpha(\Omega_i+\Omega_j)]\\ &&\times\left(\frac{\bar{n}_i+1}{ \bar{n}_i-\bar \mu+\bar \gamma\omega_i}-\frac{\bar{n}_i}{(\bar{n}_i-1)-\bar \mu+\bar \gamma\omega_i}\right),\nonumber \end{eqnarray} where $\bar n_i$ is chosen is such a way that $\bar n_i-(1-\bar\gamma \omega_i) < \bar\mu < \bar n_i+\bar\gamma \omega_i$. The MI phase corresponds to such $\bar t,\bar \mu$ that the system admits only a trivial solution (the energy has the minimum at $\psi_j=0$ for all $j$). Clearly, this occurs if and only if $\det(\bar t \mathcal{R}-\boldsymbol 1)\neq 0$, in other words, whenever \begin{equation}\label{eq:Mottbound} \bar t \max[\lambda(\mathcal{R})]<1, \end{equation} where $\lambda(\cdot)$ denotes the spectrum of a matrix. Once $\bar t$ exceeds the critical value (the condition (\ref{eq:Mottbound}) is violated) we enter a phase with at least one unstable mode that attains a non-zero value, determined by the full nonlinear dependence of the energy on $\psi_i$'s; obviously, the linear theory predicts only the instability and, formally, a value of the amplitude of the corresponding unstable mode tending to infinity. \subsection{Non-standard mean-field approach - populating unstable modes} The idea here is a simple one - we consider all modes that are unstable (i.e. these that violate the inequality (\ref{eq:Mottbound})). Note that for finite systems the spectrum of the matrix is discrete, hence only for specific values of $\bar t$ different eigenvectors of $\mathcal R$ become the solutions of Eq. (\ref{eq:self_consistent}), i.e., a fixed point of the linear map $\bar t\mathcal R$. Hence, in general it is more physical to consider the vectors $\psi$ belonging to the unstable manifold of the map $\bar t \mathcal R$, rather than the individual solutions of the self-consistency equation (\ref{eq:self_consistent}). We denote by $\mathcal Q(\bar t)$ the set of indices for the eigenvectors of $\bar t\mathcal R$ corresponding to the eigenvalues larger then one. These are the modes that we expect to be populated. Consequently, we define new modes, $a_k, a_k^{\dagger}$, corresponding to right and left eigenvectors of $\mathcal{R}$, $\psi^{(k)}, \overline{\psi}^{(k)}$, and related to the original modes as \begin{equation}\label{eq:modes} b_i=\sum_k \psi_i^{(k)}a_k, \quad b_i^{\dagger}=\sum_k \overline{\psi_i}^{(k)}a_k^{\dagger}. \end{equation} and express the initial (not decoupled) Hamiltonian (\ref{eq:Hamiltonian_3}) in terms of these operators. Our aim is to minimize the energy with respect to the population $\{n_k\}$ of the new modes. Taking the ground state in the form $\ket{g.s.}=\sum_{k\in\mathcal Q (\bar t)} 1/{\sqrt{n_k!}} a_k^{\dagger}{}^{n_k}\ket{0}$ we obtain \begin{equation}\label{eq:energy} \av{H_3}_{\ket{g.s.}}=\sum_{k\in \mathcal Q(\bar t)} n_k E_k + \sum_{k,l\in \mathcal Q(\bar t)} n_k(n_l-\delta_{k,l}) O_{kl}, \end{equation} where $\mathcal Q(\bar t)$ is the set of indices defined before, and \begin{eqnarray}\label{eq:defs} E_k&=&-\bar t \sum_{<i,j>} [1+\alpha(\Omega_i+\Omega_j)]\overline{\psi}_k^{(i)}\psi_k^{(j)}+\sum_i \bar \gamma\omega_i\overline{\psi}_k^{(i)}\psi_k^{(i)},\nonumber\\ O_{kl}&=&\frac{1}{2}\sum_i (2-\delta_{k,l})\overline{\psi}_k^{(i)} \overline{\psi}_l^{(i)} \psi_k^{(i)}\psi_l^{(i)}. \end{eqnarray} The number of particles is adjusted for each value of $\bar t$ to match the chemical potential $\bar \mu$ in the definition of $\mathcal{R}$. Knowing the population of the modes, we determine the distribution of the number of particles $n_i$ in the lattice as: \begin{equation}\label{eq:avnum} \av{n_i}=\sum_k n_k \bar{\psi}_i^{(k)}\psi_i^{(k)}. \end{equation} As long as the regions of non-zero (i.e. of absolute value greater then a threshold value) number of particles {\color{blue} } are disconnected we identify the phase as the BG. The boundary of this phase and the transition to the SF phase is given by $\bar t$ for which the sites with non-zero $n_i$ begin to percolate. This condition seems analogous to the approach of Niederle and Rieger\cite{Niederle2013}, however, there is an important difference. While in \cite{Niederle2013} the percolation of mean-field occupations of sites is directly taken as a superfluid border, in our approach we rebuild these occupations from the Hartree-Fock-like procedure discussed above. \section{Gutzwiller-ansatz}\label{sec:Gutzwiller} The results obtained following the approach discussed above will be in the next Section confronted with the standard Gutzwiller approach. In the latter the minimization of the energy functional $E[\psi]=\langle \psi | H |\psi\rangle$ over the product states of the form \begin{equation} |\psi\rangle=\prod_i \sum_n f_i(n) |n\rangle_i, \label{gucio} \end{equation} and subject to the normalization condition $\langle \psi | \psi\rangle=1$ is performed. One needs to minimize over the expansion coefficients $f_i(n)$, with $|n\rangle_i$ denoting the Fock states at site $i$. The numerical minimization of such a nonlinear problem is simple for a homogeneous system, as minimization variables $f_i(n)$ without loss of generality may be considered site-independent. For low densities one limits possible occupations to, say, $n_{max}=5$ making it a five-parameter minimization (taking into account the normalization). In such case the standard conjugate-gradient minimization algorithms always converge then to the global minimum. In contrast, in the presence of disorder the number of minimized parameters increases to $n_{max}L$ where $L$ is the number of sites. More importantly, the energy landscape of the energy functional $E[\psi]$ may contain plenty of local minima in the presence of disorder. To reach the (hopefully) global ground state additional precautions have to be made (starting from different initial conditions, perturbing the found minima to check whether they are the local or global ones etc.). We have used a 2D lattice which contained $M\times M=40\times40$ lattice sites ($L=1600$) with $n_{max}=4$. Periodic boundary conditions have been assumed. In the MI phase the mean field solution yields a Fock state at each site with a vanishing variance of the occupation number $\sigma^2_i\equiv \langle (b_i^\dagger b_i)^2 \rangle - \langle b_i^\dagger b_i \rangle^2$. Thus a convenient criterion for the disappearance of MI is that $\sigma^2 \equiv \max \sigma^2_i$ exceeds a given threshold value $s_m$. Of course the obtained results depend to some extent on the value of $s_m$. Practically the dependence is quite small, we found $s_m=0.001$ leads to an almost perfect agreement between the Gutzwiller ansatz prediction and the eigenvalue condition \eqref{eq:Mottbound}. Having the distribution of $\sigma^2_i$ one could define the BG-SF transition as a border at which non-zero $\sigma^2_i$ percolate. However, we employ another method calculating the classical property, the superfluid fraction (SF), $\rho_s$ which should vanish in BG phase. It is obtained using the ``boost'' method \cite{Lieb02,Roth03,Damski03} transferring the system to the moving frame by making the tunneling amplitude complex. Explicitly for tunneling along the $x$-axis we change $t\rightarrow t\exp(\pm i\varphi)$ for tunnelings (with sign corresponding to the direction of tunneling and $\varphi$ being a small angle). This corresponds to the presence of a constant flux proportional to $\varphi$ and the SF is obtained as \cite{Lieb02,Roth03,Damski03} $\rho_s= (E(\varphi)-E(0))/Nt\varphi^2$ where N is the total number of atoms and $E(\varphi)$ is the ground state energy at a given value of $\varphi$. In practice the parabolic dependence of $E(\varphi)$ on $\varphi$ is tested to extract a reliable SF. Let us, however, mention that the SF calculated in this manner is not entirely correct in the presence of the density dependent tunneling (for a detailed discussion of superfluid fraction definition in various situations see \cite{Rousseau2014}). We believe, however, that since in our case the density dependent tunneling terms are small, the onset of the superfluidity may be well estimated by the traditional approach. Before discussing the results let us stress that the SF fraction introduced in this way has little in common with the ``averaged SF order parameter'' criticized in \cite{Niederle2013}. The results obtained for the superfluid border using a proper superfluid fraction reproduce, in fact, the percolation border of \cite{Niederle2013} with quite a good accuracy. They are shown below together with the HF percolation threshold. \section{Results}\label{sec:results} \begin{figure}[!ht] \includegraphics[scale=0.5,trim=21 0 0 0]{Mott_1D_mu_NEW}\includegraphics[scale=0.5,trim=16 0 0 0,clip=true]{Mott_2D_mu_NEW} \includegraphics[scale=0.5,trim=21 0 0 0]{Mott_1D_corr_NEW} \includegraphics[scale=0.5,trim=16 0 0 0,clip=true]{Mott_2D_corr_NEW} \includegraphics[scale=0.5,trim=21 0 0 0]{Mott_1D_uncorr_NEW} \includegraphics[scale=0.5,trim=16 0 0 0,clip=true]{Mott_2D_uncorr_NEW} \caption{Comparison of the numerical and theoretical boundaries of the Mott lobes for the Hamiltonians a), d) $H_1$, b), e) $H_2$ and c), f) $H_3$ in 1D (first column) and 2D (second column). The numerical results for 1D were obtained for $L=100,1000,10000$ while in 2D for $L=40,50,80$. Theoretical boundaries (with shaded area) are based on the analysis of the random variables characterizing spectrum of the matrix $\mathcal{R}$ (conditions in Eqs. (\ref{eq:Mott_1D}) and (\ref{eq:Mott_2D}) in the thermodynamic limit.}\label{fig:Mott} \end{figure} Consider first the MI phase and determination of its borders (MI lobes). We study the 1D systems of different lengths $L=100,1000,10000$ and the 2D system of size $L=40,50,80$ with open boundary conditions. This allows for semi-analytical expressions for the MI borders in the thermodynamic limit. In this limit the properties of the system do not depend on the density of the impurities \cite{Mering2011pra} thus we conveniently choose $p=0.5$. We also choose $\gamma=0.3, \alpha=0.1$ in the model. Those parameters are, as mentioned above, determined by the boson-fermion interaction strength. For too strong interactions one would have to consider higher bands \cite{Mering2011pra,Jurgensen12} not included in our model. First, we find the boundary of the MI phase analyzing the spectrum of the matrix $\mathcal R$. In the 1D case $\mathcal R$ is a tridiagonal matrix with the following random off-diagonal upper ($X_i^{+}$) and lower ($X_i^{-}$) elements \begin{eqnarray}\label{eq:Xpm_def} X_i^{\pm}&=&[1+\alpha(\Omega_i+\Omega_{i\pm 1})]\\ &&\times\left(\frac{\bar{n}_i+1}{ \bar{n}_i-\mu+\bar \gamma \omega_i}-\frac{\bar{n}_i}{(\bar{n}_i-1)-\mu+\bar \gamma \omega_i}\right)\nonumber \end{eqnarray} and zeros elsewhere. As discussed by Mering and Fleishhauer \cite{Mering2008pra} (see also \cite{Buonsante2009pra}) in the thermodynamic limit the border of the fully incompressible Mott phase may be determined from {\it non-random} the situation, i.e., assuming that all sites $\omega_i$ are identical. It is due to the fact that it is the tunneling which kills MI and it is optimal at resonance, i.e., between identical sites. This argument holds also for our model with density induced tunneling terms. Then the random variables $X_i^{\pm}$ take the same value for each $i$ which allows us to estimate the spectrum of $\mathcal R$ using the formula for the tridiagonal Toeplitz matrix \cite{Bottcher2005}. \begin{equation}\label{eq:Lambda_1D} \Lambda_k=2 \sqrt{X_i^{+} X_{i}^{-}} \cos\left(\frac{k\pi}{L+1} \right),\quad k=1,\ldots,L. \end{equation} Now for large $L$ the condition (\ref{eq:Mottbound}) takes the form \begin{equation}\label{eq:Mott_1D} 2 \bar t \max\left[\sqrt{X_i^{+} X_{i}^{-}}\right]<1. \end{equation} Analogously for 2D we obtain \begin{equation}\label{eq:Mott_2D} 2 \bar t \max\left[\sqrt{X_{i,j}^{+} X_{i,j}^{-}}+\sqrt{Y_{i,j}^{+} Y_{i,j}^{-}}\right]<1, \end{equation} where $X_{i,j}^{\pm}, Y_{i,j}^{\pm}$ are given by expressions similar to (\ref{eq:Xpm_def}) (see Appendix \ref{app:RandomMatrices} for a more detailed derivation). In Fig. \ref{fig:Mott} we compare the theoretical boundary of the Mott lobes for the Hamiltonians $H_1, H_2, H_3$ with the one obtained for finite systems of sizes $L=100,1000,10000$ (1D) and $L=40,80,100$ (2D). Note that for the binary disorder Mott lobes with non-integer average filling equal to $m+(1-p), m=0,\ldots$ (number of free sites) appear between the standard Mott lobes \cite{Buonsante2009pra} as for $m<\bar \mu<m+\bar \gamma$ the local number of bosons is site-dependent, i.e., \begin{eqnarray} &&m< \bar \mu < \bar n_i < 1+\bar \mu < 1+m+\bar \gamma,\quad \text{for }\omega_i=0\\ &&m-\bar \gamma <\bar \mu-\bar\gamma < \bar n_i < 1+\bar\mu-\bar\gamma < 1+m,\quad \text{for }\omega_i=1,\nonumber \end{eqnarray} which gives $n_i=1+m$ in the first case and $n_i=m$ in the second. \begin{figure}[!ht] \includegraphics[scale=0.5]{Var_1D2D_NEW} \caption{Variance of the eigenvalues of $\mathcal R$ for 1D (dashed lines) and 2D systems (continuous lines) described by the Hamiltonians $H_1$ (orange), $H_2$ (green) and $H_3$ (purple).} \label{fig:spectra} \end{figure} In order to assess whether in any of the three analyzed cases a direct MI-SF transition is possible or if the system always passes through an intermediate BG phase we analyse the properties of the spectrum of $\mathcal R$. As indicated in Eqs. (\ref{eq:Mott_1D}) and (\ref{eq:Mott_2D}) the most relevant eigenvalues stemming from homogeneous region are $\sqrt{X_i^+X_i^-}$ for any $i$ in that region. For large enough region the cosine term in (\ref{eq:Lambda_1D}) may be approximated as unity. It turns out that for the Hamiltonian $H_1$ for each choice of the parameter $\gamma$ there exists $\mu$ for which the variance of the random variables $\sqrt{X_i^+X_i^-}$, ${\rm Var}(\Lambda)$ vanishes. This value of $\mu$ corresponds to the tip of the MI lobe. In other words, at this point, the matrix $\mathcal R$ is exactly Toeplitz and not random. We compare the variances of the largest eigenvalues of $\mathcal R$, corresponding to large homogeneous regions, for the three Hamiltonians in 1D and 2D in Fig.~\ref{fig:spectra}. Clearly, for $H_2$ and $H_3$ the variance of the eigenvalues never becomes zero. Moreover, the minimum variance of the eigenvalues for $H_3$ is much larger than for $H_2$. This leads us to the conclusion that $H_1$ may have triple points and a direct MI-SF phase transition whenever ${\rm Var}(\Lambda)$ vanishes. The addition of the disorder in the tunneling term in $H_2$ removes the triple points. In this case at least a thin layer of the BG phase surrounds the Mott lobe everywhere. In the last situation, in which the disorder in the tunneling is independent from the random chemical potential, the area of the BG should be much wider despite the same strength of the disorder. We now determine the boundary of the BG by finding the distribution of $\av{n_i}$ in the lattice as in Eq. (\ref{eq:avnum}). We compute the energy and the number of particles in the SF clusters using the redefined modes and the information about their population. The total number of particles which is adjusted to match the chemical potential provides an estimate of the condensate fraction. In Fig. \ref{fig:num_distribution} we show a typical distribution of the SF clusters in the regime of the BG and in the point of the transition to the SF phase in 2D. \begin{figure}[ht!] \includegraphics[scale=1]{BG_2D_mu035_mu_NEW}\includegraphics[scale=1]{SF_2D_mu035_mu_NEW} \includegraphics[scale=1]{BG_2D_mu035_cor_NEW}\includegraphics[scale=1]{SF_2D_mu035_cor_NEW} \includegraphics[scale=1]{BG_2D_mu035_un_NEW}\includegraphics[scale=1]{SF_2D_mu035_un_NEW} \caption{Typical distribution of the SF clusters, i.e., the regions of the lattice with non-zero local occupation by the ``superfluid'' particles, in the systems described by Hamiltonians $H_1$, $H_2$, $H_3$. In the first column the clusters do not percolate, hence the phase is identified as the Bose-glass. In the second column, the clusters begin to percolate and the SF phase emerges.}\label{fig:num_distribution} \end{figure} In Fig.~\ref{fig:glass2D} we compare the phase diagrams for the Hamiltonians $H_1, H_2, H_3$ obtained using the percolation approach with the BG-SF border coming from the Gutzwiller ansatz as explained in the previous sections. Observe that the percolation approach gives a consistently larger BG region, notably for $H_3$, although in all the cases we find that the BG regions are quite small in 2D for our choice of parameters. This is in contrast to the 1D situation with prominent BG regions \cite{Buonsante2009pra} for $H_1$ and also for $H_2$ and $H_3$ (not shown). We do not present the results for 1D since the mean field does not give reasonable quantitative predictions in this case. \begin{figure}[ht!] \includegraphics[scale=0.3]{h14040rn.png}\\ \includegraphics[scale=0.3]{h24040rn.png}\\ \includegraphics[scale=0.3]{h34040rn.png} \caption{Bose-glass regions for Hamiltonians $H_1$, $H_2$, $H_3$ (from top to bottom). Black line (with circles) give the MI-BG border, red line (triangles) corresponds to Gutzwiller ansatz BG-SF border obtained using superfluid fraction threshold, green line (squares) estimates the same border using HF percolation approach. Note that the latter is consistently higher than the SF prediction.} \label{fig:glass2D} \end{figure} \section{Summary and discussion}\label{sec:summary} We studied the Bose-Hubbard model with binary disorder obtained by pinning a secondary (fermionic) type of atoms in the optical potential. We revealed the following differences between the models describing such system: i) the model with disorder entering only in the potential term admits a direct MI-SF transition, which is not in contradiction with the theorem of inclusions; ii) in the model with disorder given by the same random variable affecting the tunneling and the chemical potential the transition to the SF phase goes always through an intermediate Bose-glass phase; iii) the disorder in the tunneling and in the potential given by independent random variables makes the intermediate Bose-glass region a bit thicker then the correlated disorder of point ii). Still the BG region in 2D for the given type and strength of disorder yields only a rather narrow slip around the Mott lobes. Moreover we introduced a new type of mean-field approach for disordered systems, which combines the local mean-field approach with a simple ``Hartree-Fock-like'' procedure. Its advantage stems from the fact that maintaining the simplicity of the local mean-field approach it allows one to bring some spatial correlations into the description. \acknowledgments The authors wish to thank R. Chhajlany and R. Augusiak for discussion. The work of O.D.\ and J.Z.\ has been supported by Polish National Science Centre within project No. DEC-2012/04/A/ST2/00088. M.L. and J.S. acknowledge the financial support from Spanish Government Grant TOQATA (FIS2008-01236), EU IP SIQS, EU STREP EQuaM, and ERC Advanced Grants QUAGATUA and OSYRIS. J.S. acknowledges the support of fundaci\'o Catalunya - La Pedrera. M.\L. acknowledges support of the Polish National Science Center by means of project no. 2013/08/T/ST2/00112 for the PhD thesis, and a research grant DEC- 2011/01/N/ST2/02549. M.\L. also acknowledges a special stipend of Smoluchowski Scientific Consortium "Matter Energy Future". Numerical simulations were performed thanks to the PL-Grid project: contract number: POIG.02.03.00-00-007/08-00 and at Deszno supercomputer (IF UJ) obtained in the framework of the Polish Innovation Economy Operational Program (POIG.02.01.00-12-023/08).
\section{Introduction} The availability of short, intense laser pulses has opened new regimes of laser-plasma interaction, rendering possible the excitation and study of various localized structures in the plasma. Amidst a plethora of excitations observed, particular role is played by electromagnetic solitons, \ie, self-trapped pulses characterized (and sustained) by a balance of dispersion or diffraction and nonlinearity. In particular, we shall be interested in so-called relativistic solitons, for which the electromagnetic field amplitude is intense enough to set plasma electrons in relativistic motion. Relativistic solitons\ have been predicted by analytical theory\rf{akhiezer56,marburger1975, kozlov79,kaw92,esirkepov1998} and simulations\rf{Bulanov_92,bulanov1999,naumova2001,Esirkepov_02,bulanov2006,wu2013}, and their signatures have been observed in experiments\rf{borghesi2002,romagnani2010,sarri2010,sarri2011,Chen_07,Pirozhkov_07,Sylla_12}. They can be thought of as electromagnetic pulses trapped in a plasma density cavitation with over-dense boundaries. For one-dimensional (1D) plasma geometry, a vast number of soliton\ families have been identified and studied\rf{esirkepov1998,Farina01a,Farina05,Farina05,Sanchez_2011a,Sanchez_2011b}, while higher-dimensional solitons\ have also been encountered in simulations\rf{bulanov2001b,Esirkepov_02}. In over-dense plasmas near the critical density, solitons\ can be excited by a long intense pulse incident on a plasma density gradient\rf{marburger1975,wu2013}. In under-dense plasmas, relativistic solitons\ have been observed behind the wake left by an intense and short pulse\rf{bulanov2001b,naumova2001,Esirkepov_02, bulanov2006}. \edit{In this case, soliton\ creation is linked to the downshift of the laser pulse frequency as it propagates through the plasma, which leads to an effective reduction of the critical density and trapping of laser pulse energy in the form of a soliton\rf{bulanov2001b}. Therefore, from this point on, we will use the term over-dense plasma in association to the frequency $\omega$ of the soliton. The latter may be lower than the frequency of the laser pulse that excited the soliton. } \edit{Simulations\rf{bulanov1999,bulanov2001b,wu2013} and experiments\rf{borghesi2002,romagnani2010} show} that multiple solitons\ may be excited by a single laser pulse, and one may expect that these solitons\ can interact with each other\rf{bulanov2001b}. In a previous publication\rf{saxena2013} we presented numerical studies of interactions of standing electromagnetic solitons\ of the form predicted in \refref{esirkepov1998}, within the relativistic cold fluid framework. For low soliton\ amplitudes, a phenomenology similar to Nonlinear Scr\"{o}dinger (NLS equation) equation soliton interaction\rf{hasegawa1995} has been observed, \eg\ involving the formation of bound states, under certain circumstances. However, soliton\ interaction for larger amplitudes departs from NLS equation\ phenomenology. The aim of the present work is to gain more insight in the origin of the NLS equation\ behavior of small amplitude soliton\ interaction in over-dense plasmas, but also to explore how deviations from NLS equation\ behavior arise. Starting from the one-dimensional relativistic cold fluid model we develop a perturbative treatment based on multiple scale analysis\rf{taniuti1969,dauxois2006, hoyle2006}, in which the small parameter is the electromagnetic field amplitude. {Under the assumptions of immobile ions and circular polarization} we derive a perturbed NLS-type equation (pNLS equation) which describes the evolution of the electromagnetic field envelope. In our expansion, localization of the soliton\ solution is introduced naturally as a result of the assumption of {a carrier frequency $\omega$} smaller but similar to the plasma frequency, {$\omega\lesssim\omega_{pe}=\sqrt{N_{0}e^2/m_e \epsilon_0}$, where $N_0$ is the plasma background density}. The dominant nonlinear term is a `focusing' cubic nonlinearity, while at higher order a `defocusing' quintic one also appears. They both result from the perturbative expansion of the relativistic $\gamma$ factor. Additional higher order nonlinear terms result from the ponderomotive coupling of the field to the plasma. To lowest order our pNLS equation\ equation reduces to a focusing NLS equation. \edit{Our study of soliton\ interactions using the fluid and pNLS equation\ models shows that soliton collisions are inelastic. This is commonly regarded as a signature of non-integrability of the governing equations\rf{fordy1990}, and is in stark contrast with the elastic collisions of solitons of the (integrable) cubic NLS\ equation. We note that although a distinction between solitary waves and solitons based on the nature of their interactions can be made\rf{scott1973}, here we adhere to the common practice in laser-plasma interaction literature of referring to any solitary wave solution as a soliton.} The NLS equation\ is ubiquitous in physics since it appears generically as an envelope equation describing propagation of weakly nonlinear waves in dispersive media\rf{hoyle2006,dauxois2006}. It has been widely used in describing phenomena such as Benjamin-Feir type modulational instabilities, solitons\rf{dauxois2006} and, more recently, rogue waves\rf{dysthe1999,veldes2013}. Interestingly, the NLS equation\ can be derived by symmetry considerations alone\rf{hoyle2006}. However, for specific applications the coefficients of the various terms must be determined through a multiple scale analysis procedure. Then, the NLS equation\ is obtained as a compatibility condition, imposed for secular term suppression, in third order in the small expansion parameter. The analogous first- and second-order compatibility conditions are also physically meaningful, as they yield the linear dispersion relation and the associated group velocity for the envelope. Higher order compatibility conditions contribute additional nonlinear terms\rf{newell1992}, which may lead to a non-integrable perturbed NLS equation. In plasma physics, use of the cubic NLS equation\ has a long history, for example to describe modulated electrostatic wavepackets\rf{kourakis2005} {and weakly relativistic laser-plasma interactions\rf{kates1989,kuehl1993,kuehl1993b}. For the case of linear polarization, in the highly under-dense plasma limit $\omega_{pe}/\omega\ll1$, fifth order terms in the multiple scale expansion have been partially retained in \refref{kuehl1993}. The main contribution of the present paper is the derivation of the fifth order terms for the over-dense, near-critical case, with circular polarization. } This paper is structured as follows: \refsect{s:model} recalls the relativistic cold fluid model that will be the starting point for this work. The multiple scale expansion is described in \refsect{s:expansion} leading to the derivation of the pNLS equation, which is our main result. In \refsect{s:nlse} we study in some detail the NLS equation\ limit of our expansion, comparing numerical results of cold fluid model simulations for soliton\ interaction to classical predictions of quasiparticle theory for soliton interaction. In \refsect{s:numerics} we compare numerical simulations of soliton\ interaction in the three levels of description: cold fluid model, pNLS equation\ and NLS equation. In \refsect{s:concl} we discuss our findings and present our conclusions. \edit{\refappe{s:ic} describes the relation between fluid and envelope initial conditions, while \refappe{s:numfluid} provides details on the numerical implementation of the fluid code.} \section{\label{s:model} Relativistic cold fluid model} Our starting point is the relativistic cold fluid plasma-Maxwell model (see, for example, \refref{gibbon2005} for the derivation and history of the model) in one spatial dimension. We assume a cold plasma of infinite extent with immobile ions that provide a neutralizing background. Considering infinite plane waves, propagating along the $x$-direction and working in the Coulomb gauge, the longitudinal and transverse component of Ampere's law are expressed as \begin{subequations}\label{eq:fluid} \begin{equation}\label{eq:longitudinal} \frac{\partial^2 \Phi}{\partial x\partial t}+\frac{N}{\gamma}P_x=0\,, \end{equation} and \begin{equation}\label{eq:wave} \frac{\partial^2\textbf{\ensuremath{\mathbf{A}_{\perp}}}}{\partial x^2}-\frac{\partial^2\textbf{\ensuremath{\mathbf{A}_{\perp}}}}{\partial t^2}=\frac{N}{\gamma}\textbf{\ensuremath{\mathbf{A}_{\perp}}}\,, \end{equation} respectively. The fluid momentum equation is \begin{equation}\label{eq:Px} \frac{\partial P_x}{\partial t}=\frac{\partial}{\partial x}\left(\Phi-\gamma\right)\,, \end{equation} while Poisson equation yields \begin{equation}\label{eq:Poisson} N=1+\frac{\partial^2\Phi}{\partial x^2}\,. \end{equation} Here, \begin{equation}\label{eq:gamma} \gamma=\sqrt{1+P_x^2+\ensuremath{\mathbf{A}_{\perp}}^2}\,, \end{equation} \end{subequations} is the relativistic factor, while $\mathbf{P}(x,t)=\mathbf{p}(x,t)-\mathbf{A}(x,t)$ is the generalized momentum of the electron fluid, $\mathbf{p}(x,t)$ the kinetic momentum normalized to $m_ec$, $N$ is the fluid charge density normalized to the plasma background density $N_0$, $\mathbf{\ensuremath{\mathbf{A}_{\perp}}}=A_y\mathbf{\hat{j}}+A_z\mathbf{\hat{k}}$ is the transverse vector potential and $\Phi$ is the scalar potential, both normalized to $m_ec^2/e$. Moreover, time and length are respectively normalized to the inverse of the plasma frequency $\omega_{p0}$ and the corresponding skin depth $c/\omega_{p0}$. In our one-dimensional modelling we have taken the longitudinal vector potential $A_x=0$, while we have used the conservation of transverse canonical momentum, assuming an initially cold plasma, to write \refeq{eq:wave} and \refeq{eq:gamma}. \section{\label{s:expansion} Multiple scale expansion} Multiple scale analysis (see, e.g. \refref{dauxois2006}) seeks to describe perturbatively a system of differential equations by assuming it evolves in different, well separated temporal and spatial scales. Our treatment follows the standard procedure of removal of secular terms by imposing suitable solvability conditions at each order in perturbation analysis. However, our treatment of the dispersion relation in \refsect{s:dispersion} is novel and has been devised in order to relate the electromagnetic field frequency to the small parameter of the expansion, while introducing localization of solutions in an over-dense plasma in a natural way. We proceed by introducing the scaled (or slow) variables $T_j = \epsilon^{j} t$ and $X_j = \epsilon^{j} x$, where $\epsilon$ is a small parameter and $j=0,1,2,\ldots\,$. We assume that the fields within the plasma are small, so that we may expand the fluid model variables as \begin{subequations}\label{eq:exp} \begin{equation}\label{eq:expAy} A_y (x,t) = \sum_{j=1}^\infty \epsilon^{j} a_j(X_0,X_1,\ldots,T_0,T_1,\ldots)\,, \end{equation} \begin{equation}\label{eq:expAz} A_z (x,t) = \sum_{j=1}^\infty \epsilon^{j} b_j(X_0,X_1,\ldots,T_0,T_1,\ldots)\,, \end{equation} \begin{equation}\label{eq:expPhi} \Phi (x,t) = \sum_{j=1}^\infty \epsilon^{j} \phi_j(X_1,\ldots,T_1,\ldots)\,, \end{equation} \begin{equation} P_x (x,t) = \sum_{j=1}^\infty \epsilon^{j} p_j(X_1,\ldots,T_1,\ldots)\,, \end{equation} \begin{equation}\label{eq:Nexp} N (x,t) = \sum_{j=0}^\infty \epsilon^{j} n_j(X_1,\ldots,T_1,\ldots)\,, \end{equation} and \begin{equation}\label{eq:gammaExp} \gamma = \sum_{j=0}^\infty \epsilon^{j} \gamma_j = 1 + \frac{1}{2}(p_1^2+a_1^2+b_1^2)\epsilon^2+\ldots\,. \end{equation} Moreover we will need \begin{equation}\label{eq:deltaExp} \delta \equiv 1/\gamma = \sum_{j=0}^\infty \epsilon^{j} \delta_j = 1 - \frac{1}{2}(p_1^2+a_1^2+b_1^2)\epsilon^2+\ldots\,. \end{equation} \end{subequations} In writing \refeq{eq:Nexp} we took into account \refeqset{eq:Poisson}{eq:expPhi}. We assumed that the longitudinal quantities $\Phi,\,P_x,\, N$ do not depend on the fast time scale $T_0$ associated with the transverse electromagnetic field oscillations, {and took into account that for CP pulses there is no generation of harmonics of the basic frequency $\omega$\rf{kates1989,kaw92}}. These assumptions are consistent with CP soliton\ solutions of \refref{esirkepov1998}, the interactions of which we shall study here. Defining $\DT{i}\equiv \partial/\partial T_i$ and $\DX{i}\equiv \partial/\partial X_i$, we get \begin{subequations}\label{eq:der} \begin{equation} \frac{\partial}{\partial t} = \DT{0} + \epsilon \DT{1} + \epsilon^2 \DT{2} + \ldots\,, \end{equation} \begin{equation} \frac{\partial}{\partial x} = \DX{0} + \epsilon \DX{1} + \epsilon^2 \DX{2} + \ldots\,, \end{equation} \begin{equation} \frac{\partial^2}{\partial x \partial t} = \DX{0}\DT{0} + \epsilon (\DX{0}\DT{1} + \DX{1}\DT{0}) + \ldots\,, \end{equation} \end{subequations} \etc We proceed by substituting \refeq{eq:exp} and \refeq{eq:der} into \refeq{eq:fluid}, and collecting terms in different orders of $\epsilon$. \subsection{Order \texorpdfstring{$\epsilon^0$}{0}} The only equation that contains terms of order $\epsilon^0$ is \refeq{eq:Poisson}, which gives \begin{equation}\label{eq:n0} n_0(X_0,X_1,\ldots,T_0,T_1,\ldots) = 1\,. \end{equation} \subsection{\label{s:dispersion} Order \texorpdfstring{$\epsilon^1$}{1}: linear dispersion relation} Collecting terms of order $\epsilon$, \refeqs{eq:longitudinal}{eq:Px} give $p_1 = n_1 = 0$, while $\phi_1$ remains unspecified. Taking into account \refeq{eq:n0}, we get for the $y$ component of \refeq{eq:wave}, \begin{equation}\label{eq:wave1} \DX{0}^2 a_1 - \DT{0}^2 a_1 = a_1\,. \end{equation} We are interested in plane wave solutions of this linear wave equation, which have the form: \begin{equation}\label{eq:a1} a_1 (X_0,X_1,\ldots,T_0,T_1,\ldots)= a(X_1,\ldots,T_1,\ldots) e^{i(k X_0 - \omega T_0)}\, +\, c.c.\,, \end{equation} where $+\, c.c.$ denotes the complex conjugate of the preceding expression. Plugging \refeq{eq:a1} back into \refeq{eq:wave1}, we see that these solutions satisfy the usual plasma dispersion relation \begin{equation}\label{eq:disp} \omega^2 = 1 + k^2\,. \end{equation} Here, we are interested in over-dense plasmas, in which $\omega<1$ and thus \begin{equation} k^2=\omega^2-1<0\,, \end{equation} \ie, $k$ is imaginary and the waves are localized. Moreover, we are interested here in describing solutions close to small amplitude solitons, which have $\omega\simeq1$. Therefore, $|k|$ is small and this fact in combination with the dispersion relation, \refeq{eq:disp}, suggests the following expansions \begin{equation}\label{eq:kwExp} \omega = \sum_{j=0}^\infty \epsilon^{j} \omega_j\,,\qquad k = \sum_{j=1}^\infty \epsilon^{j} k_j\,. \end{equation} Substituting \refeq{eq:kwExp} into the dispersion relation \refeq{eq:disp} we obtain \begin{equation}\label{eq:dispAppr} \omega = 1 - \frac{\epsilon^2}{2}|k_1|^2+\ldots\,, \end{equation} where we used $k_1=\pm i|k_1|$. This implies that expressions such as $\epsilon|k_1|\,X_0$, $\epsilon^2|k_1|^2 T_0$ \etc\ appearing in the oscillating part of \refeq{eq:a1}, do in fact represent slow variations. Thus, we may include them into the envelope, \ie, we may write \refeq{eq:a1} as \begin{equation}\label{eq:a0corr} a_1 (X_0,X_1,\ldots,T_0,T_1,\ldots) =\\ a(X_1,\ldots,T_1,\ldots) e^{-i T_0}\,+\,c.c.\,. \end{equation} Note that the dependence on $X_0$ has naturally dropped. \subsection{Order \texorpdfstring{$\epsilon^2$}{2}} Collecting terms of order $\epsilon^2$ we get from \refeq{eq:longitudinal} and \refeq{eq:Poisson} $p_2 = n_2 = 0$, while \refeq{eq:Px} yields \begin{equation}\label{eq:Px2} \DX{1}\phi_1-\frac{1}{2}\DX{0}(a_1^2+b_1^2) = 0 \,, \end{equation} respectively. For circular polarization (CP), we have $a= - i\,b$, and therefore \refeq{eq:gammaExp} yields \begin{equation} \gamma_2 = \frac{1}{2}(a_1^2+b_1^2)=2\,|a|^2\,, \end{equation} independent of $T_0$. Thus, \refeq{eq:Px2} simplifies to \begin{equation}\label{eq:DX1phi1} \DX{1}\phi_1=0\,. \end{equation} Collecting terms of order $\epsilon^2$ we get for the $y$ component of \refeq{eq:wave}, \begin{equation}\label{eq:wave2} \mathcal{L} a_2 = 2 \DT{0} \DT{1} a_1 = -2\,i\, \frac{\partial a}{\partial T_1} e^{-i T_0}\,, \end{equation} where we introduced the operator \begin{equation} \mathcal{L} \equiv (\DX{0}^2-\DT{0}^2-1)\,. \end{equation} The term on the right hand side is resonant forcing term for the linear operator $\mathcal{L}$. Therefore we impose the solvability condition \begin{equation}\label{eq:solv2} \frac{\partial a}{\partial T_1} = 0\,. \end{equation} Then \refeq{eq:wave2} becomes $\mathcal{L}a_2 = 0\,,$. It has plane wave solutions that can be included in $a_1$ in \refeq{eq:a0corr}, so we may take $a_2 = 0$. From \refeq{eq:gammaExp} we then find $\gamma_3 =0$. \subsection{Order \texorpdfstring{$\epsilon^3$}{3}: NLS equation\ limit} Collecting terms of order $\epsilon^2$ and using \refeq{eq:DX1phi1}, we get from \refeqs{eq:longitudinal}{eq:Px}, $p_3=n_3=0$, and \begin{equation}\label{eq:Px3} \DX{2}\phi_1+\DX{1}(\phi_2-2|a|^2) = 0\,, \end{equation} respectively. Operating on \refeq{eq:Px3} with $\DX{1}$ and using \refeq{eq:DX1phi1}, we obtain \begin{equation}\label{eq:Px3poisson} \DX{1}^2(\phi_2-2|a|^2) = 0\,, \end{equation} which, requiring that $\phi_2$ and $a$ vanish as $X_1\rightarrow\infty$, implies \begin{equation}\label{eq:phi2a} \DX{1}(\phi_2 -2|a|^2) = 0\,. \end{equation} Then in turn, \refeq{eq:Px3} implies \begin{equation}\label{eq:DX2phi1} \DX{2}\phi_1 = 0\,. \end{equation} Collecting terms of order $\epsilon^2$ we get for the $y$ component of \refeq{eq:wave}, \begin{align}\label{eq:wave3} \mathcal{L} a_3 & = -\DX{1}^2 a_1 + (\DT{1}^2 + 2 \DT{0} \DT{2}) a_1 - 2|a|^2 a_1\,,\\ & = -\frac{\partial^2 a}{\partial X_1^2} e^{-i T_0} - 2\,i\, \frac{\partial a}{\partial T_2} e^{-i T_0} - 2\,|a|^2 a\, e^{-i T_0}\,, \end{align} where we also used \refeq{eq:solv2}. The terms in the right hand side are all resonant with the operator $\mathcal{L}$, leading to the solvability condition \begin{equation}\label{eq:NLS} i\, \frac{\partial a}{\partial T_2} + \frac{1}{2}\frac{\partial^2 a}{\partial X_1^2} + |a|^2 a = 0\,. \end{equation} This is a {nonlinear Schr\"ondinger (NLS) equation}. Imposing the solvability condition \refeq{eq:NLS}, we get from \refeq{eq:wave3}, $\mathcal{L}a_3 = 0$, or $a_3 =0$. According to the standard multiple scale treatment of waves in the fluid plasma description, the cubic NLS equation~\refeq{eq:NLS} is obtained at this order, and the iterative expansion procedure stops here. {An exception is \refref{kuehl1993}, where the limit of small density ($\omega_{pe}\ll\omega$) has been considered and a different scaling of the slow variables has been derived, partially including terms of fifth order in $\epsilon$. The absence of harmonic terms for circular polarization makes the $\epsilon$-expansion much simpler in our case and allows us to keep all terms up to fifth order in the following, deriving a single perturbed NLS equation.} \subsection{Order \texorpdfstring{$\epsilon^4$}{4}} From \refeq{eq:longitudinal} we get, with the help of \refeq{eq:DX1phi1} and \refeq{eq:DX2phi1}, \begin{equation} \DX{1}\DT{1}\phi_2 + p_4 = 0\,. \end{equation} Operating with $\DT{1}$ on \refeq{eq:phi2a} and using \refeq{eq:solv2} we obtain \begin{equation} \DX{1}\DT{1}\phi_2 = 0\,, \end{equation} and therefore $p_4 = 0$. \refEq{eq:Poisson} gives \begin{equation}\label{eq:n4} n_4 = \DX{1}^2\phi_2\,. \end{equation} Noting that \begin{equation} \gamma_4 = -2\,|a|^4\,, \end{equation} we get from \refeq{eq:Px} \begin{equation}\label{eq:Px4} \DX{1}\phi_3 + \DX{3}\phi_1+\DX{2}(\phi_2 + 2|a|^2) =0 \,. \end{equation} Collecting terms of order $\epsilon^4$ we get for the $y$ component of \refeq{eq:wave}, \begin{align}\label{eq:wave4} \mathcal{L} a_4 & = 2\DT{0}\DT{3}a_1 - 2 \DX{1}\DX{2}a_1\\ & = -2\,i\DT{3}a\, e^{-iT_0} - 2 \DX{1}\DX{2}a\, e^{-iT_0}\,. \end{align} Requiring that resonant forcing terms vanish, we obtain the solvability condition \begin{equation}\label{eq:solv4} i\DT{3}a + \DX{1}\DX{2}a =0\,. \end{equation} Finally, \refeq{eq:wave4} implies $a_4$ may be included into $a_1$ and we write $a_4 = 0$. \subsection{Order \texorpdfstring{$\epsilon^5$}{5}} \refEq{eq:longitudinal} gives \begin{equation}\label{eq:longitudinal5} p_5 + \DX{3}\DT{1}\phi_1 + \DX{2}\DT{1}\phi_2+\DX{1}\DT{2}\phi_2+\DX{1}\DT{1}\phi_3=0\,, \end{equation} where we have used \refeq{eq:DX2phi1} and \refeq{eq:DX1phi1}. Applying $\DX{1}$ on \refeq{eq:Px4} and using \refeq{eq:DX1phi1} and \refeq{eq:phi2a} we get \begin{equation}\label{eq:phi3Lapl} \DX{1}^2\phi_3 = 0\,, \end{equation} which implies \begin{equation}\label{eq:DX1phi3} \DX{1}\phi_3 = 0\,. \end{equation} On the other hand, applying $\DT{1}$ on \refeq{eq:Px4} and using \refeq{eq:solv2} and \refeq{eq:DX1phi3} we get \begin{equation}\label{eq:DX3DT1phi1} \DX{3}\DT{1}\phi_1+\DX{2}\DT{1}\phi_2=0\,. \end{equation} Thus, we may use \refeqs{eq:DX1phi3}{eq:DX3DT1phi1} to get from \refeq{eq:longitudinal5} \begin{equation}\label{eq:p5} p_5 = -\DX{1}\DT{2}\phi_2 = -2\DT{2}\DX{1}|a|^2\,, \end{equation} where, in the last step, we have used \refeq{eq:phi2a}. From Poisson equation \refeq{eq:Poisson} we obtain \begin{equation}\label{eq:n5} n_5 = 2 \DX{2}\DX{1}\phi_2 = 4\DX{2}\DX{1}|a|^2\,. \end{equation} \refEq{eq:Px} gives \begin{equation}\label{eq:Px5} \DX{1}(\phi_4-\gamma_4)+\DX{2}\phi_3+\DX{3}(\phi_2-\gamma_2)+\DX{4}\phi_1 = 0\,. \end{equation} Operating on \refeq{eq:Px5} with $\DX{1}$ and using \refeq{eq:phi2a}, \refeq{eq:Px3poisson} and \refeq{eq:DX1phi3} we obtain \begin{equation}\label{eq:Px5poisson} \DX{1}^2(\phi_4-\gamma_4) = 0\,. \end{equation} Noting that $\delta_4=6\,|a|^4$, the wave equation gives at order $\epsilon^5$ \begin{equation}\label{eq:wave5} \mathcal{L}\, a_5 = -2\DX{1}\DX{3}a_1-\DX{2}^2 a_1 +2 \DT{0}\DT{4}a_1+\DT{2}^2 a_1 + 6 |a|^4 a_1 + n_4 a_1\,. \end{equation} All terms on the right hand side are resonant forcing terms, and with the help of \refeq{eq:n4} and \refeq{eq:Px3poisson} we are led to the solvability condition \begin{equation}\label{eq:solv5} -2\DX{1}\DX{3}a-\DX{2}^2 a -2\,i\,\DT{4}a +\DT{2}^2 a + 6 |a|^4 a + 2 a\DX{1}^2|a|^2 = 0\,. \end{equation} We note, that we do not employ a parabolic approximation but rather maintain the term involving second time derivative. Finally, \refeq{eq:wave5}, allows one to write $a_5 = 0$. \subsection{Collecting different orders: pNLS equation} The strategy we follow in order to write a perturbed NLS equation, inspired by \refref{newell1992} (Chapter 3), is to multiply the solvability conditions imposed at various orders with appropriate constants, so as to form the expansions of the differential operators \refeq{eq:der}. Specifically, we form the sum \begin{equation} 2\,i\,\epsilon\times(\ref{eq:solv2}) + 2\,\epsilon^2\times(\ref{eq:NLS}) + 2\,\epsilon^3\times(\ref{eq:solv4})- \epsilon^4\times(\ref{eq:solv5})\,, \end{equation} which, using \refeq{eq:der}, reads \begin{equation} i\frac{\partial a}{\partial t} + \frac{1}{2}\frac{\partial^2 a}{\partial x^2}+\epsilon^2 |a|a - 3\,\epsilon^4 |a|^4 a -\epsilon^2 \frac{\partial^2 |a|^2}{\partial x^2} a - \frac{1}{2} \frac{\partial^2 a}{\partial t^2}= 0 \end{equation} Rescalling, \begin{equation}\label{eq:scale} X=\epsilon x,\qquad T=\epsilon^2 t\,, \end{equation} we arrive at \begin{equation}\label{eq:env} i\frac{\partial a}{\partial T}+\frac{1}{2}\frac{\partial^2 a}{\partial X^2}+|a|^2 a = \epsilon^2\left( 3\, |a|^4 a + \frac{\partial^2|a|^2}{\partial X^2} a + \frac{1}{2}\frac{\partial^2 a}{\partial T^2}\right)\,, \end{equation} which is our main result. \refEq{eq:env} has the form of a singularly perturbed nonlinear Schr\"{o}dinger equation (pNLS equation), since the highest order time derivative appears in the perturbation term. Note that, due to our treatment of the dispersion relation in \refsect{s:dispersion}, \refeq{eq:env} is written in the lab frame, without the need of a coordinate transformation to a frame moving with the group velocity $v_g=\omega'(k)$ which is usually required in the derivation of NLS-type equations. We therefore bypass any problems related to the fact that $\omega'(k)$ as obtained from \refeq{eq:disp} is imaginary. Solution of \refeq{eq:env} for $a(X,T)$ also determines the rest of the variables in the perturbation expansion of the cold fluid model. Indeed, we get from \refeq{eq:Nexp}, \refneq{eq:n0}, \refneq{eq:n4} and \refneq{eq:n5}, in a similar manner as above, \begin{equation}\label{eq:n} N = 1+ 2\,\epsilon^4\frac{\partial^2 |a|^2}{\partial X^2}+\mathcal{O}(\epsilon^6)\,, \end{equation} \ie, $n_5$ may be included in the fourth order terms. Similarly, \refeq{eq:p5} gives \begin{equation}\label{eq:p} P = - 2\, \epsilon^5 \frac{\partial^2 |a|^2}{\partial T \partial X} + \mathcal{O}(\epsilon^6)\,, \end{equation} Finally, noting that \refeq{eq:DX1phi1} yields \begin{equation}\label{eq:phi1Lapl} \DX{1}^2\phi_1=0\,, \end{equation} and forming the sum \begin{equation} \epsilon\times(\ref{eq:phi1Lapl})+\epsilon^2\times(\ref{eq:Px3poisson})+\epsilon^3\times(\ref{eq:phi3Lapl}) +\epsilon^4\times(\ref{eq:Px5poisson})\,, \end{equation} we obtain \begin{equation}\label{eq:longBalance} \frac{\partial^2}{\partial X^2}(\Phi-\gamma) = \mathcal{O}(\epsilon^5)\,. \end{equation} Noting that $\Phi$, $\gamma$, do not depend on $X_0$, and taking into account boundary conditions at $x\rightarrow\pm\infty$, we obtain \begin{align} \Phi & = \gamma-1+\mathcal{O}(\epsilon^5)\label{eq:phigamma}\\ & = 2|a|^2\,\epsilon^2 - 2|a|^4\,\epsilon^4 +\mathcal{O}(\epsilon^5)\,. \end{align} Therefore, the solution of \refeq{eq:env} for $a(X,T)$, also determines $N,\, P$ and $\Phi$. The cubic $|a|^2 a$ and quintic terms $3\,\epsilon^2 |a|^4 a$ come from the expansion of the relativistic factor $\gamma$ and therefore correspond to relativistic corrections to the optical properties of the plasma, \ie, they pertain to transverse dynamics. The cubic term results in compression of the pulse, while the quintic term has the opposite effect (due to the difference in sign). The term $a\,\partial_{XX}|a|^2$ takes into account longitudinal effects, namely the coupling of the electromagnetic field to the plasma through charge separation caused by ponderomotive effects [see \refeq{eq:n}]. Moreover, \refeq{eq:p} shows that the balance of ponderomotive $\partial_X \gamma$ and electrostatic force $\partial_X \phi$ implied by \refeq{eq:longBalance}, is violated at order $\epsilon^5$ [see also \refeq{eq:Px}]. \section{\label{s:nlse} NLS equation\ limit} Neglecting terms of order $\epsilon^2$, \refeq{eq:env} reduces to a NLS equation\rf{sulem1999} \begin{equation}\label{eq:NLSsimple} i\, \frac{\partial a}{\partial T} + \frac{1}{2}\frac{\partial^2 a}{\partial X^2} + |a|^2 a = 0\,, \end{equation} which has both moving and standing envelope soliton solutions, which we now briefly study in order to establish connection with cold-fluid model solitons. \subsection{Moving soliton solutions} Moving soliton solutions of \refeq{eq:NLSsimple} have the form\rf{dauxois2006} \begin{equation}\label{eq:nlsMoving} a(X,T) = \alpha_0\, \mathrm{sech}\left[\alpha_0 (X-u_e\,T)\right]e^{i\,u_e\,(X-u_p\,T)}\,, \end{equation} where $u_e$ is the envelope (or group) velocity, $u_p$ is the phase velocity, and the amplitude $a_0$ is given by \begin{equation} \alpha_0=\sqrt{u_e^2-2\,u_e\,u_p}\,. \end{equation} Note that $u_e$ and $u_p$ refer to the scaled variables $X$ and $T$. We can go back to the original variables by using \refeq{eq:AyAzEnv} of \refappe{s:ic}, which gives \begin{equation} A_y = F_0(x,t)\,\cos\left[u_e\epsilon x-(1+u_e u_p \epsilon^2)t\right]\,, \end{equation} where $F_0(x,t)=2\, \alpha\, \epsilon\,\mathrm{sech}\left[\alpha_0\epsilon\left(x-\epsilon u_e t\right)\right]$. A similar expression can be written for $A_z$, while \refEq{eq:n}, \refneq{eq:p} and \refeq{eq:phigamma} provide expressions for the remaining fluid variables. We see that the phase and group velocity read \begin{equation} v_{ph} = \frac{1+u_e u_p \epsilon^2}{u_e \epsilon}\,, \end{equation} and \begin{equation} v_{g} = \epsilon u_e\,, \end{equation} respectively. Therefore, in terms of the cold fluid model, moving soliton solutions \refeq{eq:nlsMoving} represent slowly propagating solitons. We show an example of propagating NLS equation\ solitons \refeq{eq:nlsMoving} with $u_e=0.9$, $u_p=0.1$ and $\epsilon=0.141$ in \reffig{fig:099_moving}. Although this is only an approximate soliton\ solution of the pNLS equation\ and cold fluid model, we observe propagation at the predicted group velocity $\epsilon u_e$, while the solution approximately maintains its shape. An exact propagating soliton\ solution would have to be determined by methods similar to those used, for example in \refrefs{kaw92,Sanchez_2011a}. This is however outside the scope of this work which emphasizes standing soliton\ interactions. \begin{figure*}[ht!] \includegraphics[width=\textwidth, clip=true]{publ_all_1s_moving_ue0p9_up0p1_omega0p99} \caption{(color online)\label{fig:099_moving} Simulation of a propagating soliton\ with $u_e=0.9$, $u_p=0.1$ and $\epsilon=0.141$, using (a) the fluid model \refeq{eq:fluid}, (b) pNLS equation\ \refeq{eq:env}, (c) NLS equation\refneq{eq:NLS}.} \end{figure*} \subsection{Standing soliton solutions} Standing soliton solutions of \refeq{eq:NLSsimple} have the form \begin{equation}\label{eq:NLSsol} a(X,T) = \phi(X) e^{- i\lambda T}\,, \end{equation} where the frequency $\lambda$ is to be determined. A well known soliton\ family of \refEq{eq:NLSsimple} has the form (see, e.g., \refref{dauxois2006}) \begin{equation}\label{eq:sech} \phi(X,T) = \alpha\, \mathrm{sech}[\alpha(X-X_0)]\,, \end{equation} where the amplitude $\alpha$ is constant. Substituting \refeqs{eq:NLSsol}{eq:sech} into \refeq{eq:NLSsimple} we obtain \begin{equation}\label{eq:lambda} \lambda=-\frac{1}{2}\alpha^2\,. \end{equation} Using \refeq{eq:AyAzEnv} of \refappe{s:ic}, one may go back to the original cold-fluid model variables, \begin{subequations}\label{eq:sechA} \begin{equation}\label{eq:sechAy} A_y(x,t) = C_0(x)\,\cos(t-\epsilon^2\alpha^2\,t/2)\,, \end{equation} \begin{equation} A_z(x,t) = C_0(x)\,\sin(t-\epsilon^2\alpha^2\,t/2)\,, \end{equation} \end{subequations} where $C_0(x)=2\epsilon \alpha\, \mathrm{sech}[\epsilon\alpha(x-x_0)]$. Note that the term $-\epsilon^2 \alpha^2\,t/2$ corresponds to the term $-\epsilon^2 |k_1|^2\,t/2$ in \refeq{eq:dispAppr} which has been included into the envelope. This correction to $\omega$ therefore corresponds to the slow oscillations of the soliton envelope, and we may identify $\alpha= |k_1|$. A connection to exact cold-fluid mode solitons\ of Esirkepov~\etal\rf{esirkepov1998} which have \begin{equation}\label{eq:Esirkepov0} A_y= R(x)\, \cos(\omega t) \,,\qquad A_z= R(x)\, \sin(\omega t)\,, \end{equation} where \begin{equation}\label{eq:R} R(x) = \frac{2\sqrt{1-\omega^2}\mathrm{cosh}\left[\sqrt{1-\omega^2}(x-x_0)\right]} {\mathrm{cosh}^2\left[\sqrt{1-\omega^2}(x-x_0)\right]+\omega^2-1} \end{equation} can now be established. Taking into account that \refeq{eq:dispAppr} yields \begin{equation}\label{eq:omegaepsk1} \omega^2=1-\epsilon^2\,|k_{1}|^2, \end{equation} $A_y$ from \refeq{eq:Esirkepov0} becomes, in leading order in $\epsilon$, \begin{equation}\label{eq:EsirkepovSmall} A_y= 2\epsilon|k_{1}|\mathrm{sech}\left[\epsilon|k_{1}|(x-x_0)\right]\cos(t-\epsilon^2\,|k_{1}|^2\,t/2)\,. \end{equation} Comparison of \refeq{eq:sechAy} with \refeq{eq:EsirkepovSmall}, once again shows that we may identify $\alpha=|k_{1}|$. The value of $|k_1|$ still remains unspecified, \edit{apart from the requirement that $|k_1|\sim\mathcal{O}(1)$ which follows from \refeq{eq:kwExp}}. For the study of a single soliton, or interacting solitons\ of the same amplitude, we may, without loss of generality, set $|k_{1}|=1$, \ie, work with $\alpha=1$ in \refeq{eq:sechA}. Then, the small parameter $\epsilon$ in the expansion is related to $\omega$ through \begin{equation}\label{eq:omegaeps} \epsilon=\sqrt{1-\omega^2}\,. \end{equation} However, the indeterminancy of $k_1$ in \refeq{eq:EsirkepovSmall} allows to model interactions of solitons\ with different frequences (and therefore amplitudes) using a single small parameter $\epsilon$, see \refsect{s:numerics}. \subsection{\label{s:quasiparticle} Quasiparticle approach to soliton interactions} Soliton interactions of NLS equation\ \refneq{eq:NLSsimple} have been studied through different methods and are well understood, particularly when the separation of the two solitons is large, see for example \refref{hasegawa1995}. In that case, one may consider the solitons as 'quasiparticles', \ie, independent entities that exert a force to one another and to a good approximation maintain their shape during their interaction\rf{karpman1981,gordon1983,hasegawa1995} {(adiabatic approximation)}. One then may approximate the evolution of the solitons with a set of few coupled ordinary differential equations for the evolution of the soliton parameters. A particularly useful perspective is that of Gordon\rf{gordon1983}, who shows that within this framework the dynamics of two solitons may be described by the system of ODEs \begin{subequations}\label{eq:gordon} \begin{align} \ddot{D} &=-8\,e^{-D}\, \cos(\Psi)\,,\\ \ddot{\Psi} &= 8\,e^{-D}\, \sin(\Psi)\,, \end{align} \end{subequations} where $D(T)$ and $\Psi(T)$ are the soliton peak-to-peak distance and relative phase, respectively, $1\pm\dot{\Psi}/2$ are their amplitudes and a dot indicates derivative with respect to $T$. Therefore, the solitons exert to each other an effective force with magnitude that decreases exponentially with distance $D$ and sign that depends on their relative phase $\Psi$. For the particular case of solitons of equal initial amplitude, $\dot{\Psi}_0=0$, and zero initial phase difference, we see from \refeq{eq:gordon} that the force is attractive, while the relative phase remains zero, leading to the formation of an oscillatory bound state. A detailed calculation\rf{gordon1983,hasegawa1995} shows that the peak-to-peak separation varies as \begin{equation}\label{eq:DX} D(T) = D_0 + 2\ln|\cos(2\,e^{-D_0/2} \, T)|\,, \end{equation} where $D_0\gg1$ is the initial distance. \refEq{eq:DX} implies that the time after which the solitons collide for the first time is\rf{gordon1983,hasegawa1995} \begin{equation}\label{eq:Tcoll} T_{coll}=\frac{1}{2}e^{D_0/2}\cos^{-1}\left(e^{-D_0/2}\right) \simeq \frac{\pi}{4}e^{D_0/2}\,. \end{equation} We note that in scaled variables collision time $T_{coll}$ is, according to \refeq{eq:Tcoll}, independent of $\omega$. However, when we go back to units of $\omega_{pe}^{-1}$ and $c/\omega_{pe}$ for time and space, respectively, through \refeqset{eq:scale}{eq:omegaeps}, we obtain \begin{equation}\label{eq:tc} t_{\mathrm{coll}}=\frac{\pi}{4(1-\omega^2)}e^{\sqrt{(1-\omega^2)}\,d_0/2}\,. \end{equation} where $t_{\mathrm{coll}}=T_{coll}/\epsilon^2$ and $d_0=D_0/\epsilon$. For large enough $d_0$, $t_{\mathrm{coll}}$ decreases for increasing $\omega$. Analytical predictions of \refeq{eq:tc} is compared with numerical simulations in \refsect{s:numerics}. For solitons of equal initial amplitude, $\dot{\Psi}_0=0$ and initial relative phase $\Psi_0$, \edit{the seperation versus time reads\rf{hasegawa1995} \begin{equation}\label{eq:DwPhase} D(T) = D_0 + \ln\left|\frac{\mathrm{cosh}(-\kappa_1\,T)+\cos(\kappa_2\,T)}{2}\right|\,, \end{equation} where \[ \kappa_1=4\,e^{-D_0/2}\,\sin\left(\Psi_0/2\right),\qquad \kappa_2=4\,e^{-D_0/2}\cos\left(\Psi_0/2\right)\,. \] Thus, the solitons eventually drift apart since, as we see from \refeq{eq:gordon}, the force is not attractive for all $T$. For large $T$, after the solitons drift apart, the soliton amplitudes differ by \begin{equation}\label{eq:Da} |\Delta a| = |a_1-a_2| = |4\,e^{-D_0/2}\cos\left(\Psi_0/2\right)|\,. \end{equation} } The minimum peak-to-peak distance is \begin{equation}\label{eq:Dmin} D_{min} = D_0+\ln\left[\frac{1}{2}\left(\mathrm{cosh}\left(-\pi\tan\frac{\Psi_0}{2}\right) -1\right)\right] \end{equation} reached at time \begin{equation} T_{min}=\frac{\pi}{4\cos(\Psi_0/2)}e^{D_0/2} \end{equation} or, in units of $\omega_{pe}^{-1}$, \begin{equation} t_{\mathrm{min}}=\frac{\pi\,e^{\sqrt{(1-\omega^2)}\,d_0/2}}{4(1-\omega^2)\cos(\Psi_0/2)}\,. \end{equation} From \refeq{eq:Dmin} one finds\rf{hasegawa1995} that the collision is inhibited for initial phase $\Psi_0>\Psi_c$, where \begin{equation} \tan\frac{\Psi_c}{2} = \frac{1}{\pi} \mathrm{cosh}^{-1}\left(1 + 2\, e^{-D_0}\right)\,. \end{equation} In the case of solitons with different initial amplitudes, $\dot{\Psi}_0 \neq 0$ and no initial relative phase, $\Psi_0=0$, one finds that the solitons form an oscillatory bound state, with their distance oscillating periodically between a minimum and a maximum distance. Qualitative results can also be obtained for this case, and may be found in \refref{hasegawa1995}. \section{\label{s:numerics} Numerical simulations} This section compares numerical simulations of soliton\ interactions of the three different levels of description: the cold fluid model \refeq{eq:fluid}, the pNLS equation\ \refeq{eq:env}, and the NLS equation\ \refneq{eq:NLS}. Where appropriate, comparisons of the simulation results and the quasiparticle predictions of \refsect{s:quasiparticle} are also presented. We investigate the interactions of standing solitons\ first found by Esirkepov\etal\rf{esirkepov1998}. Labeling each soliton\ by an index $(j)$, we have for the cold fluid model variables: \begin{widetext} \begin{subequations}\label{eq:Esirkepov} \begin{equation}\label{eq:EsirkepovA} A_y^{(j)}(x,t)= \frac{2\sqrt{1-\omega_j^2}\mathrm{cosh}^2(\zeta_j)} {\mathrm{cosh}^2(\zeta_j)+\omega_j^2-1} \cos(\omega_j\,t+\theta_j)\,,\qquad A_z^{(j)}(x,t)= \frac{2\sqrt{1-\omega_j^2}\mathrm{cosh}^2(\zeta_j)} {\mathrm{cosh}^2(\zeta_j)+\omega_j^2-1} \sin(\omega_j\,t+\theta_j) \,, \end{equation} \begin{equation} E_x^{(j)} (x,t) = \frac{4\,(1-\omega_j^2)^{3/2}\,\mathrm{cosh}(\zeta_j)\sinh(\zeta_j)}{\left(\mathrm{cosh}^2(\zeta_j)+\omega_j^2-1\right)^2}\,,\qquad E_y^{(j)} (x,t) = \omega_j\, A_z^{(j)}\,,\quad E_z^{(j)} = -\omega_j\, A_y^{(j)}\,, \end{equation} \begin{equation} N^{(j)}(x,t) = 1 + \left(1-\omega_j^2\right)^2\,\frac{\mathrm{cosh}(4\zeta_j)-2\,(2\,\omega_j^2-1)\,\mathrm{cosh}(2\,\zeta_j)-3}{\left(\mathrm{cosh}^2(\zeta_j)-1+\omega_j^2\right)^3}\,,\quad P_x =0\,, \end{equation} \end{subequations} \end{widetext} where $\zeta_j=\sqrt{1-\omega_j^2}(x-x_j)$ and $\theta_j$ is an initial phase. The soliton amplitude is $R_0^{(j)}\equiv|\mathbf{A_{\perp}}(x_j)|=2\sqrt{1-\omega^2}/\omega^2$. It takes its maximum value $R_{0,\mathrm{max}}=\sqrt{3}$ for $\omega_{\mathrm{min}}=\sqrt{2/3}$, where the branch of Esirkepov solitons\ terminates because the minimum local density vanishes. In the following, we study interactions of pairs of solitons\ with frequences $\omega_1$, $\omega_2$ and initial phase difference $\Psi_0=\theta_2-\theta_1$, \ie, with initial conditions given by $A_y(x,0)=A_y^{(1)}(x,0)+A_y^{(2)}(x,0)$, \etc\ For the density we take care that the boundary condition $N(x,0)\rightarrow1$ as $x\rightarrow\pm\infty$ is satisfied by using the initial condition $N(x,0)=N^{(1)}(x,0)+N^{(2)}(x,0)-1$. For the simulations using pNLS equation\ \refeq{eq:env}, we introduce a parameter $k_1^{(j)}$ for each soliton. We then set $|k_1^{(1)}|=1$, which fixes $\epsilon=\sqrt{1-\omega_1^2}$ through \refeq{eq:omegaepsk1}, and introduce the soliton\ amplitude as $\alpha_j=\sqrt{1-\omega_j^2}/\epsilon$, effectively using the freedom to choose $k_1^{(j)}=\alpha_j$. Then, according to \refeqset{eq:arai0}{eq:aprapi} of \refappe{s:ic}, initial conditions for the soliton\ centered at $x_j=X_j/\epsilon$ are given as \begin{subequations}\label{eq:envInit} \begin{equation}\label{eq:aInit} a^{(j)}(X,0)= \frac{\alpha_j\mathrm{cosh}^2(\alpha_j(X-X_j))} {\mathrm{cosh}^2\left[\alpha_j(X-X_j)\right]+\omega_j^2-1}\, e^{-i\theta_j}\,, \end{equation} \begin{equation}\label{eq:NLSinit} \left.\frac{\partial a^{(j)}}{\partial T}\right|_{T=0} = \frac{i\,\alpha_j^2}{2}\,a^{(j)}(X,0)\,. \end{equation} \end{subequations} Finally, for the simulations with the NLS equation\ \refeq{eq:NLSsimple}, we expand \refeq{eq:aInit} to lowest order in $\epsilon$, to obtain the NLS equation\ soliton \begin{equation} a^{(j)}(X,0)= \alpha_j\mathrm{sech}\left[\alpha_j(X-X_j)\right]\, e^{-i\theta_j}\,. \end{equation} All simulations are carried out with the package \texttt{XMDS2}\rf{dennis2013}, using the pseudospectral method with Fourier space evaluation of partial derivatives and a fourth order Adaptive Runge-Kutta-Fehlberg scheme (known as ARK45) for time-stepping. More details on the implementation are provided in \refappe{s:numfluid}. We validated our fluid code by verifying that a single soliton\ with $\omega_1=0.98$, \ie, of the largest amplitude $R_0^{(j)}=0.414$ studied here, follows \refeq{eq:Esirkepov} up to $t=2\times10^{4}$. Initial conditions \refeq{eq:envInit} do not correspond to an exact soliton\ solution of the pNLS equation. {However, by integrating such initial conditions up to $t=T/\epsilon^2=2\times10^4$, we show in \reffig{f:publ_env_1s_err} that the relative error introduced by the choice of initial conditions, remains small for the maximum amplitude considered in the following numerical examples (corresponding to $\omega\geq0.98$)}. \begin{figure}[htb!] \begin{center} \includegraphics[width=\cwidth]{publ_env_1s_err} \end{center} \caption{(color online)\label{f:publ_env_1s_err} Estimate of the relative error introduced by neglecting higher order terms in pNLS equation~\refeq{eq:env}. We plot $\mathrm{max}\left[a(x,T)-a(x,0)\right]/a(0,0)$ for solutions of \refeq{eq:env} up to $T/\epsilon^2=2\times10^4$, with initial condition \refeq{eq:envInit} with $x_1=0$. Dots correspond to $\omega=(0.999,\, 0.99,\, 0.98,\, 0.97,\, 0.96,\, 0.95)$. The solid line is the best fit to the data, showing that $\mathrm{rel.\ error}\sim \epsilon^{4.3}$. } \end{figure} \subsection{Small amplitude limit} For small amplitude solitons\ we find excellent agreement between cubic NLS equation, pNLS equation\ and cold-fluid model predictions. \subsubsection{Solitary waves of equal amplitude and no phase difference} \begin{figure*}[htb!] \includegraphics[width=\textwidth, clip=true]{publ_all_2s_omega0p999_d100} \caption{(color online)\label{f:0999_d100} Comparison of simulations of two soliton\ interaction with frequencies $\omega_1=\omega_2=0.999$, $R_0^{(1)}=R_0^{(2)}\simeq0.0896$ (corresponding to $\epsilon\simeq0.0447$, $k_1^{(1)}=k_1^{(2)}=1$) and initial distance $d_0=100$ using (a) the fluid model \refeq{eq:fluid}, (b) pNLS equation\ \refeq{eq:env}, (c) NLS equation\ \refneq{eq:NLS}.} \end{figure*} A typical case in which we have formation of a bound state of solitons\ is shown in \reffig{f:0999_d100}. In these simulations, both solitons\ have frequencies $\omega_1=\omega_2=0.999$, $R_0^{(1)}=R_0^{(2)}\simeq0.0896$ (corresponding to $\epsilon\simeq0.0447$, $k_1^{(1)}=k_1^{(2)}=1$), initial distance is $d_0=100$ and there is no initial phase difference. The behavior of the exact cold fluid model solutions is captured correctly by both the NLS equation\ and pNLS equation\ models. Therefore, at very small amplitudes, the behavior may be completely understood in terms of the NLS equation. The major role in soliton attraction and in bound state formation is thus played by the leading term of the relativistic nonlinearity, \ie, by the cubic term in the NLS equation. The predictions of quasiparticle theory of \refsect{s:quasiparticle} for the collision time $t_{\mathrm{coll}}$ or the period of oscillations $t_p=2\,t_{\mathrm{coll}}$ are in excellent agreement with fluid model simulations, see \reffig{f:Tc}. \begin{figure}[htb!] \begin{center} \includegraphics[width=\cwidth]{publ_tcol_d} \end{center} \caption{(color online)\label{f:Tc} Comparison of results of quasiparticle theory for $\omega=0.999$ (red, dashed line) and $\omega=0.99$ (blue, solid line) with numerical results from simulations of the fluid model (circles and cross signs, respectively) for the first collision time $t_{\mathrm{coll}}$ for solitons\ of equal amplitude and no initial phase difference, as a function of initial separation $d_0$.} \end{figure} \subsubsection{Solitary waves of equal amplitude and finite phase difference} Next, we study the case of two solitons\ of equal frequency $\omega_1=\omega_2=0.999$ and amplitude $R_0^{(1)}=R_0^{(2)} \simeq 0.0896$ (corresponding to $\epsilon\simeq0.0447$, $k_1^{(1)}=k_1^{(2)}=1$) and a finite initial phase difference {$\Psi_0=0.1\pi$}, see \reffig{f:0999_dth01}. Also in this case, we find excellent quantitative agreement between cold fluid model, NLS equation\ and pNLS equation\ simulations. The solitons\ attract, but do not collide, in agreement with quasiparticle theory which predicts that collisions are inhibited for {$\Psi_0 > \Psi_c=0.136$}. The solitons\ reach a minimum distance $d_{\mathrm{min}}\simeq57$ at time $t_{\mathrm{min}} \simeq 3485$ and subsequently drift apart. Quasiparticle theory underestimates the minimum distance $d_{\mathrm{min}}=38.2$, while the time of minimum approach $t_{\mathrm{min}}=3720$ is in good agreement with simulations. \edit{According to quasiparticle theory, \refeq{eq:Da}, the solitons differ in amplitude by $|\Delta A|= 2\,\epsilon\, |\Delta a| = 0.037$. This is in excellent agreement with the NLS\ simulation, for which the soliton amplitudes differ at $t=10^5$ by $|\Delta A| = 0.036$. } \begin{figure*}[htb!] \includegraphics[width=\textwidth, clip=true]{publ_all_2s_omega0p999_dth_0p1pi_d100} \caption{(color online)\label{f:0999_dth01} Comparison of simulations of two soliton\ interaction with frequency $\omega_1=\omega_2=0.999$, amplitude $R_0^{(1)}=R_0^{(2)} \simeq0.0896$ (corresponding to $\epsilon\simeq0.0447$, $k_1^{(1)}=k_1^{(2)}=1$) initial distance $d_0=100$ and phase difference $\Psi_0=0.1\pi$ using (a) the fluid model \refeq{eq:fluid}, (b) pNLS equation\ \refeq{eq:env}, (c) NLS equation\ \refneq{eq:NLS}.} \end{figure*} \subsubsection{Solitary waves of unequal amplitude and no phase difference} For solitons\ of unequal frequences $\omega_1=0.999$ and $\omega_2=0.998$, and therefore also unequal amplitudes $R_0^{(1)}\simeq0.0896$, $R_0^{(2)}\simeq0.127$ (corresponding to $\epsilon\simeq0.0447$, $k_1^{(1)}=1$, $k_1^{(2)}\simeq1.4139$), with no initial phase difference, as shown for example in \reffig{f:0999_0998_d100}, the solitons\ interact and form a periodic bound state. However, their separation remains finite. Again, there is excellent agreement between all three levels of description. This behavior can be also understood in terms of quasiparticle theory of NLS equation, see \refref{hasegawa1995}. \begin{figure*}[htb!] \includegraphics[width=\textwidth, clip=true]{publ_all_2s_omega0p999_0p998_d100} \caption{(color online)\label{f:0999_0998_d100} Comparison of simulations of two soliton\ interaction with $\omega_1=0.999$, $R_0^{(1)}\simeq0.0896$ and $\omega_2=0.998$, $R_0^{(2)}\simeq0.127$ (corresponding to $\epsilon\simeq0.0447$, $k_1^{(1)}=1$, $k_1^{(2)}\simeq1.4139$) and initial distance $d_0=100$ using (a) the fluid model \refeq{eq:fluid}, (b) pNLS equation\ \refeq{eq:env}, (c) NLS equation\ \refneq{eq:NLS}.} \end{figure*} \subsection{\label{s:intense} Larger amplitudes} For solitons\ of moderately large amplitude, but still in the perturbative regime $\epsilon\ll1$, cold-fluid model simulations deviate from cubic NLS equation\ predictions and the fifth order terms need to be taken into account. \subsubsection{Solitary waves of equal amplitude and no phase difference} \refFig{f:099_d60} shows the interaction of two solitons\ of equal amplitude with $\omega_1=\omega_2=0.99$ ($R_{0,1}=R_{0,2}\simeq0.2879$, $\epsilon\simeq0.1411$, $k_1^{(1)}=k_1^{(2)}=1$), separated by $d_0=60$ and no initial phase difference. The solitons\ collide approximately at $t_{\mathrm{coll}}\simeq2800$, in good agreement with quasiparticle theory prediction $t_{\mathrm{coll}}\simeq2717$, and they form a bound state. However, contrary to the prediction of the cubic NLS equation\ model, collisions are inelastic and part of the soliton\ field decays away each time the solitons\ collide, see \reffig{f:099_d60_profiles}. The bound state is reminiscent of a system of damped oscillators: after each encounter, the solitons\ decrease in amplitude and come closer together. In turn, this implies that the period of bound state oscillations becomes shorter, as predicted by \refeq{eq:tc}. \begin{figure*}[ht!] \includegraphics[width=\textwidth, clip=true]{publ_all_2s_omega0p99_d60} \caption{(color online)\label{f:099_d60} Comparison of simulations of soliton\ interaction with frequencies $\omega_1=\omega_2=0.99$, amplitude $R_0^{(1)}=R_0^{(2)} \simeq0.2879$ ($\epsilon\simeq0.1411$, $k_1^{(1)}=k_1^{(2)}=1$) and initial distance $d_0=60$ using (a) the fluid model \refeq{eq:fluid}, (b) pNLS equation\ \refeq{eq:env}, (c) NLS equation\ \refneq{eq:NLS}.} \end{figure*} This behavior is also captured qualitatively by the pNLS equation, see \reffig{f:099_d60}(b). However, the pNLS equation\ overestimates the loss of soliton\ field at collisions and, correspondingly, the period of oscillations decreases with a faster rate than in the fluid model simulations. These deviations may be attributed to the breakdown, at the moment of collision, of the assumptions of small field magnitude and slow spatio-temporal evolution. In particular, close to the collision, the spatial scale for variations in the envelope is a few $c/\omega_{pe}$, indicating that we should expect quantitative discrepancies between the fluid and pNLS equation\ simulations. The cubic NLS equation\ approximation on the other hand, presents qualitative difference in this case. Since cubic NLS equation\ is an integrable model, all soliton collisions are elastic, leading to an exactly periodic bound state of solitons. Therefore, the inclusion of the fifth order terms in our analysis is necessary for a qualitatively correct description of soliton\ interactions. The collision time on the other hand, as predicted by quasiparticle theory of NLS equation, agrees very well with fluid simulation results even in this larger amplitude case, see \reffig{f:Tc}. The reason for this is that attraction of solitons\ is determined by their overlap. For large initial separation of the solitons\, this overlap occurs at small amplitudes. Moreover, in the limit $x\rightarrow\pm\infty$ the exact soliton\ envelope \refeq{eq:aInit} takes the form of the NLS equation\ soliton \refeq{eq:NLSinit}. Therefore, NLS equation\ approximation is a valid one in order to determine the initial attraction phase of two well separated solitons\, even for larger amplitudes. \begin{figure}[ht!] \includegraphics[width=\cwidth, clip=true]{publ_fluid_2s_omega0p99_d60_profiles} \caption{(color online)\label{f:099_d60_profiles} Two snapshots from cold fluid model simulations of soliton\ interaction with frequencies $\omega_1=\omega_2=0.99$, amplitude $R_0^{(1)}=R_0^{(2)} \simeq0.2879$ ($\epsilon\simeq0.1411$, $k_1^{(1)}=k_1^{(2)}=1$) and initial distance $d_0=60$ [see also \reffig{f:099_d60}(a)]. Dashed (blue) line corresponds to $t=0$; solid (green) line corresponds to $t\simeq10745$, \ie, at maximum separation after the third collision. \edit{The inset corresponds to the area in the gray box.} } \end{figure} \subsubsection{Solitary waves of equal amplitude and finite phase difference} As an example, we show the interaction of two solitons\ with $\omega_1=\omega_2=0.99$ ($R_{0,1}=R_{0,2}\simeq0.2879$, $\epsilon\simeq0.1411$, $k_1^{(1)}=k_1^{(2)}=1$) with initial distance $d_0=60$ and phase difference $\Psi_0=0.1\,\pi$, in \reffig{f:099_dth01}. The fluid simulations show that the solitons\ initially approach and then separate, moving in opposite directions with velocities $v_L=-0.002$ and $v_R=0.0021$ for the left and right moving soliton, respectively. The pNLS equation\ simulations agree with the fluid model predictions quantitatively, faithfully capturing the velocities of the outgoing solitons\ to be $v_L=-0.0024$ and $v_R=0.0025$. On the other hand, NLS equation\ simulations exhibit oscillations in soliton position before the latter drift apart, a feature not present in the fluid simulations. Furthermore, the velocities of the outgoing solitons\ in the NLS equation\ simulations are $v_L=-0.0005$ and $v_R=0.0007$, rather small compared to the fluid simulations. Quasiparticle predictions for the distance of minimum approach $d_{\mathrm{min}}$ and corresponding time $t_{\mathrm{min}}$ are compared with the results of the numerical simulations in \reftab{tab:mindist}, showing excellent agreement. \begin{figure*}[htb!] \includegraphics[width=\textwidth, clip=true]{publ_all_2s_omega0p99_dth_0p1pi_d60} \caption{(color online)\label{f:099_dth01} Comparison of simulations of two soliton\ interaction with frequency $\omega_1=\omega_2=0.99$ (amplitude $R_0^{(1)}=R_0^{(2)} \simeq0.2879$, $\epsilon\simeq0.1411$, $k_1^{(1)}=k_1^{(2)}=1$) initial distance $d_0=60$ and phase difference $\Psi_0=0.1\pi$ using (a) the fluid model \refeq{eq:fluid}, (b) pNLS equation\ \refeq{eq:env}, (c) NLS equation\ \refneq{eq:NLS}.} \end{figure*} \begin{table}[ht!] \begin{tabular}{l|c|c} & $d_{\mathrm{min}}\, [c/\omega_{pe}]$ & $t_{\mathrm{min}}$ $[\omega_{pe}^{-1}]$ \\ \hline quasiparticle (analytical) & 40.5 & 2751.4 \\ fluid simulation & 39.8 & 2703.0 \\ pNLS equation\ simulation & 39.7 & 2682.2 \\ NLS equation\ simulation & 40.0 & 2688.4 \end{tabular} \caption{\label{tab:mindist} Comparison of analytical prediction based on quasiparticle theory and results of numerical simulations of the different models studied here, for the distance of minimum approach $d_{\mathrm{min}}$ and corresponding time $t_{\mathrm{min}}$, for two solitons\ of frequency $\omega_1=\omega_2=0.99$, amplitude $R_0^{(1)}=R_0^{(2)} \simeq0.2879$ ($\epsilon\simeq0.1411$, $k_1^{(1)}=k_1^{(2)}=1$), initial distance $d_0=60$ and initial phase difference $\Psi_0=0.1\pi$ (see \reffig{f:099_dth01}). } \end{table} We also present the interaction of two solitons\ with $\omega_1=\omega_2=0.99$ ($R_{0,1}=R_{0,2}\simeq0.2879$, $\epsilon\simeq0.1411$, $k_1^{(1)}=k_1^{(2)}=1$), initial distance $d_0=60$ and very small phase difference $\Psi_0=10^{-6}$ in \reffig{f:099_dth1e-6}. In this case, $\Psi_0<\Psi_c$, and indeed the two solitons\ do collide. However, after a subsequent recollision, they diverge and move away from each other. The pNLS equation\ simulations capture this feature, even though the velocities of the escaping solitons\ are much larger than in the fluid simulations. However, in the NLS equation\ simulations we see that the solitons keep recolliding for many iterations. \edit{Quasiparticle theory predicts that even for small initial phase difference the solitons will eventually drift apart; however, from \refeq{eq:DwPhase} we find that the time scale required for this to happen for such a small initial phase difference is of the order of $10^{9}$, well beyond the maximum integration time $t=2\times10^5$ for which we could simulate the system.} \begin{figure*}[htb!] \includegraphics[width=\textwidth, clip=true]{publ_all_2s_omega0p99_dth_1e-6_d60} \caption{(color online)\label{f:099_dth1e-6} Comparison of simulations of two soliton\ interaction with frequency $\omega_1=\omega_2=0.99$, amplitude $R_0^{(1)}=R_0^{(2)} \simeq0.2879$ ($\epsilon\simeq0.1411$, $k_1^{(1)}=k_1^{(2)}=1$) initial distance $d_0=60$ and phase difference {$\Psi_0=10^{-6}$} using (a) the fluid model \refeq{eq:fluid}, (b) pNLS equation\ \refeq{eq:env}, (c) NLS equation\ \refneq{eq:NLS}.} \end{figure*} \edit{The cold fluid model conserves the normalized energy, $E=E_l+E_p+E_e$, where \begin{align} E_l &= \frac{1}{2}\int \left[\left(\frac{\partial A_y}{\partial t}\right)^2+\left(\frac{\partial A_y}{\partial x}\right)^2\right]dx\,,\\ E_p &= \frac{1}{2}\int \left(\frac{\partial \phi}{\partial x}\right)^2dx\,,\\ E_e &= \int \left(\gamma-1\right)n\ dx\,, \end{align} are electromagnetic, electrostatic and kinetic energy contributions, respectively. In soliton collisions with a finite phase difference, energy can be transfered between the two solitons. We show this in \reffig{f:energyExchange}, where we plot as a function of time the total energy $E_{\mathrm{tot}}$ in the computational domain and the energy $E_L$ ($E_R$) in the left (right) half of the domain, for the cold fluid model simulation of \reffig{f:099_dth1e-6}(a). We find that after the solitons separate, energy has been transfered to the soliton in the right half of the domain, while its amplitude has increased.} \begin{figure}[ht!] \includegraphics[width=\cwidth, clip=true]{publ_all_2s_omega0p99_dth_1e-6_d60_Energy} \caption{(color online)\label{f:energyExchange} Total energy $E_{\mathrm{tot}}$ in the computational domain (red, dot-dashed line) and energy $E_L$ (blue, solid line) and $E_R$ (green, dashed line) in the left and right half of the domain, respectively, corresponding to cold fluid model simulation of \reffig{f:099_dth1e-6}(a). } \end{figure} \edit{The results of \reffig{f:099_dth1e-6} indicate that for larger amplitudes, cold fluid model bound states do not persist under perturbations involving phase difference of the two solitons, leading to qualitative difference from the NLS equation\ model dynamics. \reffig{f:099_dth} illustrates that the number of re-collisions between solitons decreases as $\Psi_0$ increases. Such dynamics has been associated with chaotic scattering of solitons in the context of a perturbed NLS equation\rf{dmitriev2002}. However, determining parameters for which chaotic scattering occurs in our problem is beyond the scope of this work. \reffig{f:099_dth} suggests that truncation error in simulations can lead to the breaking of a bound state, even when there is no initial phase difference, as we have observed in \refref{saxena2013}. Therefore, we have taken care that simulations presented here are well resolved by checking that varying the spatial resolution does not affect the results. } \begin{figure*}[ht!] \includegraphics[width=\textwidth, clip=true]{publ_fluid_2s_omega0p99_dth_d60} \caption{(color online)\label{f:099_dth} Cold fluid model simulations of two soliton\ interaction with frequency $\omega_1=\omega_2=0.99$, amplitude $R_0^{(1)}=R_0^{(2)} \simeq0.2879$ ($\epsilon\simeq0.1411$, $k_1^{(1)}=k_1^{(2)}=1$) initial distance $d_0=60$ and phase difference using (a) {$\Psi_0=10^{-9}$} (b) {$\Psi_0=10^{-7}$}, (c) {$\Psi_0=10^{-5}$}.} \end{figure*} \subsubsection{Solitary waves of unequal amplitude and no phase difference} \begin{figure*}[htb!] \includegraphics[width=\textwidth, clip=true]{publ_all_2s_omega0p99_0p98_d30} \caption{(color online)\label{f:099_098_d30} Comparison of simulations of two soliton\ interaction with frequencies $\omega_1=0.99$, $\omega_2=0.98$, amplitudes $R_0^{(1)}\simeq0.2879$, $R_0^{(2)}\simeq 0.4144$ (corresponding to $\epsilon\simeq0.1411$, $k_1^{(1)}=1$, $k_1^{(2)}\simeq1.4107$), initial distance $d_0=30$ and relative phase {$\Psi_0=0$} using (a) the fluid model \refeq{eq:fluid}, (b) pNLS equation\ \refeq{eq:env}, (c) NLS equation\ \refneq{eq:NLS}.} \end{figure*} The case of solitons\ of frequencies $\omega_1=0.99$, $\omega_2=0.98$, amplitudes $R_0^{(1)}\simeq0.2879$, $R_0^{(2)}\simeq 0.4144$ (corresponding to $\epsilon\simeq0.1411$, $k_1^{(1)}=1$, $k_1^{(2)}\simeq1.4107$), initial distance $d_0=30$ and relative phase {$\Psi_0=0$}, is shown in \reffig{f:099_098_d30}. In this case, the solitons\ after some oscillations quickly diverge from each other. Simulations using the pNLS equation\ faithfully capture this behavior. On the contrary, NLS equation\ simulations show the formation of a bound state of oscillating solitons. We can therefore conclude that once again we have qualitative differences between cold-fluid model and NLS equation\ soliton\ interactions, with bound states appearing unstable within the former model. \subsection{\label{s:cqnlse} Cubic-quintic NLS equation\ approximation} Finally, we show that the qualitatively new features present in the larger amplitude simulations, may be captured by keeping only the quintic nonlinearity in \refeq{eq:env}, \ie, by the cubic-quintic NLS equation: \begin{equation}\label{eq:cqnlse} i\frac{\partial a}{\partial T}+\frac{1}{2}\frac{\partial^2 a}{\partial X^2}+|a|^2 a - 3\,\epsilon^2 \, |a|^4 a = 0\,, \end{equation} We present three different examples of soliton\ interaction dynamics under \refeq{eq:cqnlse} in \reffig{f:cqnlse}. In all cases, there is agreement at a qualitative level with cold fluid model simulations of \refeq{eq:cqnlse} of \refsect{s:intense}. This suggests that the 'defocusing' quintic term induces the qualitative changes in soliton\ interactions. However, obtaining better quantitative agreement requires keeping all terms in \refeq{eq:env}, as in \refsect{s:intense}. \begin{figure*}[hbt!] \includegraphics[width=\textwidth, clip=true]{publ_cqnls_2s} \caption{(color online)\label{f:cqnlse} Solitary wave interaction using the cubic-quintic NLS equation\ approximation \refeq{eq:cqnlse} for two solitons\ with (a) frequencies $\omega_1=\omega_2=0.99$, amplitudes $R_0^{(1)}=R_0^{(2)} \simeq0.288$ and initial distance $d_0=60$, as in \reffig{f:099_d60} , (b) frequencies $\omega_1=\omega_2=0.99$, amplitudes $R_0^{(1)}=R_0^{(2)} \simeq0.288$ initial distance $d_0=60$ and phase difference {$\Psi_0=10^{-6}$}, as in \reffig{f:099_dth01} and (c) frequencies $\omega_1=0.99$, $\omega_2=0.98$ ($R_0^{(1)}\simeq0.288$ and $R_0^{(2)}\simeq 0.414$) and initial distance $d_0=30$, as in \reffig{f:099_098_d30}. } \end{figure*} \section{\label{s:concl} Discussion and Conclusions} We studied weakly relativistic bright solitons\ and their interactions using a perturbative, multiple scale analysis and direct numerical simulations. We derived an equation for the evolution of the field envelope valid for small amplitudes, keeping terms up to order five in the small parameter. Localization of the soliton\ solutions appears naturally in our scheme through the requirement $\omega\lesssim\omega_{pe}$ obtained through the linear dispersion relation. The lowest order nonlinear effect is due to the relativistic nonlinearity, leading to a classical cubic NLS equation. The cubic nonlinearity balances pulse dispersion, leading to the formation of solitons. However, higher order terms, most importantly the quintic term resulting from the expansion of the relativistic nonlinearity, become essential at larger amplitudes. In particular, the response of the plasma to the ponderomotive force and the formation of a density cavitation which confines the soliton, is only captured by keeping fifth order terms in the small parameter, leading to the pNLS equation\ \refeq{eq:env}. At even higher amplitudes, soliton\ width becomes comparable to the wavelength \edit{of the carrier electromagnetic wave} and the perturbative description breaks down. We have demonstrated the utility of the NLS equation\ and pNLS\ equations derived here by applying them to the problem of standing solitary wave interaction. We have found that the lowest order, NLS equation\ approximation works very well for lower amplitudes, but gives qualitatively different results at higher amplitudes. For well separated solitons the quasiparticle approach provides analytical estimates for the first collision time and minimum distance of approach of two solitons. We have found that these estimates are in very good agreement with fluid simulations, even for larger amplitudes. The reason for this is that the overlapping part of well separated relativistic solitons\ can be well approximated by the tails of NLS equation\ solitons. The effect of higher order terms that leads, \eg, to inelastic collisions, only becomes significant after the solitons\ have approached each other. Once the solitons\ are sufficiently close to each other, the higher order terms become important, and lead to qualitatively different results than in NLS equation. For example, since the NLS equation\ is a completely integrable equation, its soliton collisions are elastic. However, in our fluid simulations we have clear signatures of inelastic soliton\ collisions accompanied by emission of radiation. These features are captured qualitatively by keeping the higher order terms in the pNLS equation. We can attribute the emission of radiation in collisions to the role of the 'defocusing' fifth order nonlinearity. In cases in which the total amplitude of the field remains in the perturbative regime, good quantitative agreement is obtained between the cold-fluid and pNLS equation\ simulations. {Our study suggests that the collision time for well separated solitons\ can become larger than the typical response time of the ions $\left(m_i/m_e\right)^{1/2}\,2\pi\,/\omega_{pe}$\rf{naumova2001}, and ion dynamics would have to be taken into account in future studies.} \edit{As we have seen in \reffig{f:energyExchange}, one feature of soliton collisions is the transfer of energy from one soliton to the other during the interaction. Moreover, it can be seen in \reffig{f:099_dth1e-6}, that the wave on the right hand side has a larger amplitude after the collision.} This is interesting in connection to wave-breaking of solitons, which occurs above a certain amplitude threshold, and has been proposed in the past as a means to accelerate particles \cite{esirkepov1998,Farina01a}. Our simulations indicate that soliton\ interaction is a good candidate to trigger wave-breaking, through which electromagnetic energy of the wave could be transfered to the particles. The introduction of NLS equation as the lowest order approximation to the problem of relativistic solitary interaction allowed the application of quasiparticle theory of NLS equation\ solitons in order to obtain analytical estimates for collision time and minimum distance of approach. The development of the pNLS equation\ framework, on the other hand, suggests the possibility to use further \edit{mathematical tools such as the inverse scattering transform (IST)~\rf{hasegawa1995}. For instance, in the context of Alfven waves, IST applied to the derivative-NLS equation\ was used to explain the collapse of the bright Alfven soliton and the formation of robust magnetic holes \rf{Sanchez2010a,Sanchez2010b}. The model derived here suggests that the IST of the NLS equation\ could be used in a similar manner to analyze simulations and get insight on the interaction and disappearance of solitons\ in laser-plasma interaction.} {In summary, a perturbed NLS equation describing electromagnetic envelope evolution for weakly relativistic pulses in plasmas has been derived. Auxiliary equations describe the plasma density, momentum and electrostatic potential in terms of the electromagnetic field. The pNLS model agrees very well with fluid model simulations of soliton\ interactions. Our simulations suggest that in the small but finite amplitude regime the defocusing quintic nonlinearity becomes important and soliton\ collisions are inelastic. } \section{Acknowledgements} ES would like to thank S. Skupin and F. Maucher for helpful suggestions. Work by GS was supported by the Ministerio de Ciencia e Innovaci\'{o}n of Spain (Grant No ENE2011-28489). We would like to thank the anonymous referee for many perspicacious observations.
\section{Introduction} \label{sec:intro} The study of simultaneous rational approximation to points in $\mathbb{R}^n$ started with the basic estimates of Dirichlet in \cite{Di1842}, and got full impetus in the years 1926-1938 through the transference theorems of Khintchine \cite{Kh1926a,Kh1926b} and J\'arnik \cite{Ja1938}. Around the same period of time, Mahler introduced new tools in geometry of numbers and applied them to these questions \cite{Ma1937}. In 1967, Schmidt \cite{Sc1967} enlarged the scope of the problem by studying how a fixed vector subspace $A$ of $\mathbb{R}^n$ or $\mathbb{C}^n$ can be approximated, in the ambient space, by vector subspaces of a given dimension, defined over a fixed number field $K$. For this purpose, he introduced the notion of height of such a subspace and was lead to study several angles of approximation depending on the dimension of $A$. This important work was recently revisited by Laurent \cite{La2009b} in the case where $A$ is a one-dimensional subspace of $\mathbb{R}^n$ and where $K=\mathbb{Q}$ is the field of rational numbers. Then, $A$ is spanned by a single vector $\mathbf{u}$ and there is only one angle of approximation to consider. This lead Laurent to introduce a family of $n-1$ (ordinary) exponents of approximation to points in $\mathbb{R}^n$ (interpolating between the two classical ones) and to recast the results of Schmidt in that setting through a series of inequalities relating these exponents. The purpose of this paper is to show that these inequalities describe the full spectrum of these exponents, thus answering a question of Laurent in \cite{La2009b}. The proof uses the parametric geometry of numbers introduced by Schmidt and Summerer in a series of recent papers \cite{SS2009,SS2013a}, together with the complements from \cite{R2015}. In the next section, we recall the definition of these exponents as well as the inequalities of Schmidt and Laurent which link them, and we state our main result. In Section \ref{sec:link}, we express these exponents in the context of parametric geometry of numbers. In Section \ref{sec:gen}, we introduce the notion of generalized $n$-system and, using \cite{R2015}, we reduce the proof of our main result to a combinatorial problem involving such systems. We study a particular family of generalized $n$-systems in Section \ref{sec:fam} and use them in Section \ref{sec:proof} to complete the proof. \section{Notation and main result} For each integer $n\ge 1$, we view $\mathbb{R}^n$ as an Euclidean space for the standard inner product of two vectors $\mathbf{x}$ and $\mathbf{y}$ denoted $\mathbf{x}\cdot\mathbf{y}$. We also view its exterior algebra ${\textstyle{\bigwedge}} \mathbb{R}^n = \oplus_{k=0}^n{\textstyle{\bigwedge}}^k\mathbb{R}^n$ as an Euclidean space characterized by the property that, for each orthonormal basis $\{\mathbf{e}_1,\dots,\mathbf{e}_n\}$ of $\mathbb{R}^n$, the products $\mathbf{e}_{i_1}\wedge\cdots\wedge\mathbf{e}_{i_k}$ with $0\le k\le n$ and $1\le i_1<\cdots<i_k\le n$ form an orthonormal basis of ${\textstyle{\bigwedge}} \mathbb{R}^n$. For each $k=1,\dots,n$, we also denote by ${\textstyle{\bigwedge}}^k \mathbb{Z}^n$ the lattice of ${\textstyle{\bigwedge}}^k \mathbb{R}^n$ spanned by the products $\mathbf{x}_1\wedge\cdots\wedge\mathbf{x}_k$ with $\mathbf{x}_1,\dots,\mathbf{x}_k\in\mathbb{Z}^n$. For $k=0$, we have ${\textstyle{\bigwedge}}^k \mathbb{R}^n=\mathbb{R}$ and we set ${\textstyle{\bigwedge}}^0 \mathbb{Z}^n=\mathbb{Z}$. We say that a vector subspace $S$ of $\mathbb{R}^n$ is defined over $\mathbb{Q}$ if it is generated over $\mathbb{R}$ by elements of $\mathbb{Q}^n$. For such a subspace $S$, a basis $\{\mathbf{x}_1,\dots,\mathbf{x}_k\}$ of $S\cap\mathbb{Z}^n$ as a $\mathbb{Z}$-module is also a basis of $S$ as a vector space over $\mathbb{R}$ and, following Schmidt \cite{Sc1967}, we define the \emph{height} of $S$ by \[ H(S) = \|\mathbf{x}_1\wedge\cdots\wedge\mathbf{x}_k\|, \] this being independent of the choice of $\{\mathbf{x}_1,\dots,\mathbf{x}_k\}$. Given a non-zero point $\mathbf{u}$ in $\mathbb{R}^n$ and a non-zero subspace $S$ of $\mathbb{R}^n$, we define the \emph{projective distance} between $\mathbf{u}$ and $S$ by \[ \mathrm{dist}(\mathbf{u},S) = \frac{\|\mathbf{u}\wedge\mathbf{x}_1\wedge\cdots\wedge\mathbf{x}_k\|} {\|\mathbf{u}\|\,\|\mathbf{x}_1\wedge\cdots\wedge\mathbf{x}_k\|}, \] where $\{\mathbf{x}_1,\dots,\mathbf{x}_k\}$ is any basis of $S$ over $\mathbb{R}$. Again this is independent of the choice of the basis. It is also given by \[ \mathrm{dist}(\mathbf{u},S) = \frac{\|\mathrm{proj}_{S^\perp}(\mathbf{u})\|}{\|\mathbf{u}\|} \] where $\mathrm{proj}_{S^\perp}(\mathbf{u})$ denotes the orthogonal projection of $\mathbf{u}$ on the orthogonal complement $S^\perp$ of $S$. Geometrically, it represents the sine of the smallest angle between $\mathbf{u}$ and a non-zero vector of $S$. In the work of Schmidt \cite{Sc1967}, this quantity is denoted $\psi_1(A,S)$ or $\omega_1(A,S)$ where $A=\mathbb{R}\mathbf{u}$ is the line spanned by $\mathbf{u}$. \begin{definition} Let $n\ge 1$ and let $\mathbf{u}\in\mathbb{R}^{n+1}\setminus\{0\}$. For each $j=0,\dots,n-1$, we denote by $\omega_j(\mathbf{u})$ (resp.\ $\hat{\omega}_j(\mathbf{u})$) the supremum of all real numbers $\omega$ such that, for arbitrarily large values of $Q$ (resp.\ for all sufficiently large values of $Q$), there exists a vector subspace $S$ of $\mathbb{R}^{n+1}$, defined over $\mathbb{Q}$, of dimension $j+1$, with \[ H(S)\le Q \quad\mbox{and}\quad H(S)\mathrm{dist}(\mathbf{u},S) \le Q^{-\omega}. \] \end{definition} In particular, we have $\omega_j(\mathbf{u})=\infty$ if $\mathbf{u}$ belongs to a subspace $S$ of $\mathbb{R}^{n+1}$ defined over $\mathbb{Q}$ of dimension $j+1$. Otherwise, $\omega_j(\mathbf{u})$ is the supremum of all real numbers $\omega$ for which there exist infinitely many subspaces $S$ of $\mathbb{R}^{n+1}$, defined over $\mathbb{Q}$, of dimension $j+1$ with \[ \mathrm{dist}(\mathbf{u},S) \le H(S)^{-\omega-1}. \] Theorem 13 of \cite{Sc1967} shows that \begin{equation} \label{results:eq:homega} \omega_j(\mathbf{u}) \ge \hat{\omega}_j(\mathbf{u}) \ge \frac{j+1}{n-j} \quad (0\le j\le n-1). \end{equation} Thus $\omega_0(\mathbf{u}),\dots,\omega_{n-1}(\mathbf{u})$ are non-negative although possibly infinite. The following result provides further inequalities relating these exponents of approximation. \begin{theorem}[Schmidt, Laurent] \label{results:thm:SL} Let $n\in\mathbb{N}^*$. For any non-zero $\mathbf{u}\in\mathbb{R}^{n+1}$, we have $\omega_0(\mathbf{u})\ge 1/n$ and \begin{equation} \label{results:thm:SL:eq} \frac{j\omega_j(\mathbf{u})}{\omega_j(\mathbf{u})+j+1} \le \omega_{j-1}(\mathbf{u}) \le \frac{(n-j)\omega_j(\mathbf{u})-1}{n-j+1} \quad (1\le j\le n-1), \end{equation} with the convention that the left-most ratio is equal to $j$ if $\omega_j(\mathbf{u})=\infty$. \end{theorem} The left inequality in \eqref{results:thm:SL:eq} follows from the Going-up-theorem of Schmidt \cite[Theorem 9]{Sc1967} while the right inequality follows from his Going-down-theorem \cite[Theorem 10]{Sc1967}. This was observed by Laurent in \cite{La2009b} who also introduced the exponents $\omega_j(\mathbf{u})$ for that purpose. In the same paper, Laurent also notes that each individual inequality in \eqref{results:thm:SL:eq} is best possible because their combination yields Khinchine's transference inequalities which are known to be best possible. He also deduces that, for each fixed $j\in\{0,\dots,n-1\}$, the spectrum of $\omega_j$, namely the set of all possible values $\omega_j(\mathbf{u})$, is the full interval $[(j+1)/(n-j),\infty]$. Although he restricts to points $\mathbf{u}$ with $\mathbb{Q}$-linearly independent coordinates, this is also true over the the set of all $\mathbf{u}\neq 0$ in view of \eqref{results:eq:homega}. An independent proof for the right inequality in \eqref{results:thm:SL:eq} is given by Laurent in \cite{La2009b} and, for both inequalities, by Bugeaud and Laurent in \cite{BL2010}, again for points $\mathbf{u}$ with $\mathbb{Q}$-linearly independent coordinates. Our main result below answers a question of Laurent in \cite{La2009b} by showing that the inequalities of Theorem \ref{results:thm:SL} describe the set of all possible values for the $n$-tuples $(\omega_0(\mathbf{u}),\dots,\omega_{n-1}(\mathbf{u}))$. \begin{theorem} \label{results:thm:main} Let $n\in\mathbb{N}^*$. For any $\omega_0,\dots,\omega_{n-1} \in [0,\infty]$ satisfying $\omega_0\ge 1/n$ and \begin{equation} \label{results:thm:main:eq} \frac{j\omega_j}{\omega_j+j+1} \le \omega_{j-1} \le \frac{(n-j)\omega_j-1}{n-j+1} \quad (1\le j\le n-1), \end{equation} there exists a point $\mathbf{u}\in\mathbb{R}^{n+1}$ with $\mathbb{Q}$-linearly independent coordinates such that \[ \omega_j(\mathbf{u})=\omega_j \quad\mbox{and}\quad \hat{\omega}_j(\mathbf{u})=\frac{j+1}{n-j} \quad (0\le j\le n-1). \] \end{theorem} A description of the spectrum of the $2n$ exponents $(\omega_0,\dots,\omega_{n-1},\hat{\omega}_0,\dots,\hat{\omega}_{n-1})$ was achieved by Laurent in \cite{La2009} for $n=2$ but, for larger values of $n$, the question is open. In \cite[\S1]{SS2013a}, Schmidt and Summerer propose a different larger set of $2n+2$ exponents of approximation to non-zero points in $\mathbb{R}^n$. Their spectrum also is unknown, even for $n=2$. \section{Link with parametric geometry of numbers} \label{sec:link} Fix $n\in\mathbb{N}^*$ and $\mathbf{u}\in\mathbb{R}^{n+1}\setminus\{0\}$. For each real number $Q\ge 1$, we form the convex body \[ {\mathcal{C}}_\mathbf{u}(Q) = \{\mathbf{x}\in\mathbb{R}^{n+1} \,;\, \|\mathbf{x}\|\le 1, \ |\mathbf{x}\cdot\mathbf{u}| \le Q^{-1}\} \] and, for $j=1,\dots,n+1$, we denote by $\lambda_j\big({\mathcal{C}}_\mathbf{u}(Q)\big)$ its $j$-th minimum, namely the smallest real number $\lambda>0$ such that $\lambda {\mathcal{C}}_\mathbf{u}(Q)$ contains at least $j$ linearly independent points of $\mathbb{Z}^{n+1}$. On the model of Schmidt and Summerer in \cite{SS2013a}, we define \[ L_{\mathbf{u},j}(q) = \log \lambda_j({\mathcal{C}}_\mathbf{u}(e^q)) \quad (q\ge 0,\ 1\le j\le n+1), \] and we form the map $\mathbf{L}_\mathbf{u}\colon[0,\infty)\to \mathbb{R}^{n+1}$ given by \[ \mathbf{L}_\mathbf{u}(q) = (L_{\mathbf{u},1}(q),\dots,L_{\mathbf{u},n+1}(q)) \quad (q\ge 0). \] For each $j=1,\dots,n+1$, we also define \[ {\ubar{\psi}}_{j}(\mathbf{u}) = \liminf_{q\to\infty} \frac{L_{\mathbf{u},1}(q)+\cdots+L_{\mathbf{u},j}(q)}{q} \quad\mbox{and}\quad {\bar{\psi}}_{j}(\mathbf{u}) = \limsup_{q\to\infty} \frac{L_{\mathbf{u},1}(q)+\cdots+L_{\mathbf{u},j}(q)}{q}. \] The following result connects these quantities to those from the previous section. \begin{proposition} \label{link:prop:omega} Let $n$ and $\mathbf{u}$ be as above. For each $j\in\{0,\dots,n-1\}$, we have \[ \omega_j(\mathbf{u})=\frac{1}{{\ubar{\psi}}_{n-j}(\mathbf{u})}-1 \quad\mbox{and}\quad \hat{\omega}_j(\mathbf{u})=\frac{1}{{\bar{\psi}}_{n-j}(\mathbf{u})}-1. \] \end{proposition} The proof relies on the following alternative definition for the exponents $\omega_j(\mathbf{u})$ and $\hat{\omega}_j(\mathbf{u})$. \begin{lemma}[Bugeaud-Laurent] \label{link:lemma:BL} Let $j\in\{0,\dots,n-1\}$. Then $\omega_j(\mathbf{u})$ (resp.\ $\hat{\omega}_j(\mathbf{u})$) is the supremum of all real numbers $\omega$ such that the inequalities \begin{equation} \label{link:lemma:BL:eq} \|\mathbf{z}\|\le Q \quad\mbox{and}\quad \|\mathbf{z}\wedge\mathbf{u}\|\le Q^{-\omega} \end{equation} have a non-zero solution $\mathbf{z}\in{\textstyle{\bigwedge}}^{j+1}\mathbb{Z}^{n+1}$ for arbitrarily large values of $Q$ (resp.\ for all sufficiently large values of $Q$). \end{lemma} This is proved in \cite[\S4]{BL2010} for $\omega_j(\mathbf{u})$ in the case where $\mathbf{u}$ has $\mathbb{Q}$-linearly independent coordinates but the same argument extends with little change to any non-zero vector $\mathbf{u}$ and applies also to $\hat{\omega}_j(\mathbf{u})$. \begin{proof}[Proof of Proposition \ref{link:prop:omega}] For each $j=0,\dots,n-1$ and each $Q\ge 1$, define \[ {\mathcal{K}}^{(j+1)}_\mathbf{u}(Q) = \{\mathbf{z}\in{\textstyle{\bigwedge}}^{j+1}\mathbb{R}^{n+1} \,;\, \|\mathbf{z}\|\le Q, \ \|\mathbf{z}\wedge\mathbf{u}\|\le 1 \}. \] According to \cite[\S4, Lemma 3]{BL2010}, ${\mathcal{K}}^{(j+1)}_\mathbf{u}(Q)$ is comparable to the $(j+1)$-th compound body of ${\mathcal{K}}^{(1)}_\mathbf{u}(Q)$. As the latter is in turn comparable to the dual of ${\mathcal{C}}_\mathbf{u}(Q)$ and as $\mathrm{vol}({\mathcal{C}}_\mathbf{u}(Q))\asymp Q^{-1}$, it follows that the first minimum of ${\mathcal{K}}^{(j+1)}_\mathbf{u}(Q)$ with respect to ${\textstyle{\bigwedge}}^{j+1}\mathbb{Z}^{n+1}$ satisfies \begin{align*} \lambda_1\big({\mathcal{K}}^{(j+1)}_\mathbf{u}(Q)\big) &\asymp \lambda_1\big({\mathcal{K}}^{(1)}_\mathbf{u}(Q)\big)\cdots \lambda_{j+1}\big({\mathcal{K}}^{(1)}_\mathbf{u}(Q)\big) \\ &\asymp \lambda_{n+1}\big({\mathcal{C}}_\mathbf{u}(Q)\big)^{-1}\cdots \lambda_{n-j+1}\big({\mathcal{C}}_\mathbf{u}(Q)\big)^{-1} \\ &\asymp Q^{-1} \lambda_1\big({\mathcal{C}}_\mathbf{u}(Q)\big) \cdots\lambda_{n-j}\big({\mathcal{C}}_\mathbf{u}(Q)\big) \qquad (0\le j\le n-1,\ Q\ge 1), \end{align*} with implied constants depending only on $n$ and $\mathbf{u}$ (see \cite[\S\S14--15]{GL1987}). Now the convex body of ${\textstyle{\bigwedge}}^{j+1}\mathbb{R}^{n+1}$ defined by the conditions \eqref{link:lemma:BL:eq} is $Q^{-\omega}{\mathcal{K}}_\mathbf{u}^{(j+1)}(Q^{\omega+1})$, so its first minimum with respect to ${\textstyle{\bigwedge}}^{j+1}\mathbb{Z}^{n+1}$ is bounded below and above by products of \begin{equation} \label{link:prop:omega:eq} Q^{-1} \lambda_1\big({\mathcal{C}}_\mathbf{u}(Q^{\omega+1})\big)\cdots \lambda_{n-j}\big({\mathcal{C}}_\mathbf{u}(Q^{\omega+1})\big) \end{equation} with positive constants that are independent of $Q$ and of $\omega$, when $Q\ge 1$ and $\omega+1\ge 0$. By Lemma \ref{link:lemma:BL}, $\omega_j(\mathbf{u})$ (resp.\ $\hat{\omega}_j(\mathbf{u})$) is the supremum of all real numbers $\omega$ for which this first minimum is $\le 1$ for arbitrarily large values of $Q$ (resp.\ for all sufficiently large values of $Q$). Since $\omega_j(\mathbf{u})\ge\hat{\omega}_j(\mathbf{u})\ge 0$ by \eqref{results:eq:homega}, we may restrict to values of $\omega$ with $\omega>-1$ and thus $\omega_j(\mathbf{u})$ (resp.\ $\hat{\omega}_j(\mathbf{u})$) is also the supremum of all real numbers $\omega>-1$ for which the product in \eqref{link:prop:omega:eq} is $\le 1$ for arbitrarily large values of $Q$ (resp.\ for all sufficiently large values of $Q$). However, upon writing $Q=e^{q/(\omega+1)}$ with $q>0$, the condition that this product is $\le 1$ amounts to \[ \frac{L_{\mathbf{u},1}(q)+\cdots+L_{\mathbf{u},n-j}(q)}{q} \le \frac{1}{\omega+1}, \] and the conclusion follows. \end{proof} \section{Generalized $n$-systems} \label{sec:gen} Fix an integer $n\ge 2$. In \cite[\S3]{SS2013a}, Schmidt and Summerer propose the notion of $(n,0)$-system as a model for the general behavior of the maps $\mathbf{L}_\mathbf{u}\colon[0,\infty)\to\mathbb{R}^n$ attached to non-zero points $\mathbf{u}\in\mathbb{R}^n$. The following is a version of this adapted to the present context. \begin{definition} \label{gen:def1} Let $I$ be a subinterval of $[0,\infty)$ with non-empty interior. An $n$-system on $I$ is a continuous piecewise linear map $\mathbf{P}=(P_1,\dots,P_n)\colon I\to\mathbb{R}^n$ with the following properties. \begin{itemize} \item[(S1)] For each $q\in I$, we have $0\le P_1(q)\le\cdots\le P_n(q)$ and $P_1(q)+\cdots+P_n(q)=q$. \item[(S2)] If $H$ is a non-empty open subinterval of $I$ on which $\mathbf{P}$ is differentiable, then there is an integer $r$ with $1\le r\le n$ such that $P_r$ has slope $1$ on $H$ while the other components $P_j$ of $\mathbf{P}$ with $j\neq r$ are constant on $H$. \item[(S3)] If $q$ is an interior point of $I$ at which $\mathbf{P}$ is not differentiable and if the integers $r$ and $s$ for which $P'_r(q^-)=P'_s(q^+)=1$ satisfy $r<s$, then we have $P_{r}(q)=P_{r+1}(q)=\cdots=P_{s}(q)$. \end{itemize} \end{definition} Here, the condition that $\mathbf{P}\colon I\to \mathbb{R}^n$ is piecewise linear means that the set $D$ of points of $I$ at which $\mathbf{P}$ is not differentiable (including the boundary points of $I$ that lie in $I$) is a discrete subset of $I$, and that the derivative of $\mathbf{P}$ is locally constant on $I\setminus D$. Such a map admits a left derivative $\mathbf{P}'(q^-)$ at each $q\in I$ with $q\neq \inf I$, and a right derivative $\mathbf{P}'(q^+)$ at each $q\in I$ with $q\neq \sup I$. The slope of a component $P_r$ of $\mathbf{P}$ on an open subinterval $H$ of $I\setminus D$ means the constant value of its derivative on $H$ or equivalently the slope of its graph over $H$. Figure \ref{gen:fig:S3} shows the combined graph of the functions $P_r,\dots,P_s$ over a neighborhood of $q$ under the hypotheses of Condition (S3). In that case, the functions $P_{r+1},\dots,P_{s}$ coincide to the left of $q$, while $P_{r},\dots,P_{s-1}$ coincide to its right. \begin{figure}[h] \begin{tikzpicture}[xscale=0.6,yscale=0.3] \node[draw,circle,inner sep=1pt,fill] at (6,4) {}; \draw[dashed] (6,4)--(6,1) node[below]{$q$}; \draw[thick] (3,1) -- (9,7); \draw[thick] (3,4) -- (9,4); \node[left] at (3,1) {$P_{r}$}; \node[left] at (3,4) {$P_{r+1}=\cdots=P_{s}$}; \node[right] at (9,4) {$P_{r}=\cdots=P_{s-1}$}; \node[right] at (9,7) {$P_{s}$}; \end{tikzpicture} \caption{Illustration for Condition (S3).} \label{gen:fig:S3} \end{figure} For an interval $I$ of the form $[q_0,\infty)$ with $q_0\ge 0$, the above notion of an $n$-system on $I$ is the same as that of an $(n,0)$-system on $I$ according to Definition 2.8 of \cite{R2015}. These $n$-systems have the following approximation property. \begin{theorem} \label{gen:thm:nsys} For each non-zero point $\mathbf{u}\in\mathbb{R}^n$, there exists $q_0\ge 0$ and an $n$-system $\mathbf{P}$ on $[q_0,\infty)$ such that $\mathbf{L}_\mathbf{u}-\mathbf{P}$ is bounded on $[q_0,\infty)$. Conversely, for each $n$-system $\mathbf{P}$ on an interval $[q_0,\infty)$ with $q_0\ge 0$, there exists a non-zero point $\mathbf{u}\in\mathbb{R}^n$ such that $\mathbf{L}_\mathbf{u}-\mathbf{P}$ is bounded on $[q_0,\infty)$. \end{theorem} The first assertion follows from \cite[Theorem 1.3]{R2015} which states the same approximation property for a more restricted class of maps called \emph{rigid $n$-systems}, while its converse follows from \cite[Theorem 8.1]{R2015}. The fact that we deal with non-zero points $\mathbf{u}$ of $\mathbb{R}^n$ instead of unit vectors like in \cite{R2015} does not matter for such a qualitative statement. The goal of this section is to show that this approximation property extends to the larger class of generalized $n$-systems which we will define below. In view of Proposition \ref{link:prop:omega}, this reduces the determination of the spectrum of any subsequence of $(\omega_0,\dots,\omega_{n-1},\hat{\omega}_0,\dots,\hat{\omega}_{n-1})$ over $\mathbb{R}^{n+1}\setminus\{0\}$ to a problem about generalized $(n+1)$-systems. The next two lemmas provide examples of $n$-systems. \begin{lemma} \label{gen:lemma1} Let $a,b\in\mathbb{R}$ with $0\le a<b$. There exists an $n$-system $\mathbf{P}\colon[a,b]\to\mathbb{R}^n$ such that $\mathbf{P}(a)=(a/n,\dots,a/n)$ and $\mathbf{P}(b)=(b/n,\dots,b/n)$. \end{lemma} \begin{proof} Set $q_i=(n-i)a/n+ib/n$ for $i=0,\dots,n$ and, for each $j=1,\dots,n$, define $P_j$ to be the unique continuous piecewise linear function on $[q_0,q_n]=[a,b]$ which is constant equal to $a/n$ on $[q_0,q_{n-j}]$, has slope $1$ on $[q_{n-j},q_{n-j+1}]$, and is constant equal to $b/n$ on $[q_{n-j+1},q_n]$. Then the map $\mathbf{P}=(P_1,\dots,P_n)\colon[a,b]\to\mathbb{R}^n$ is an $n$-system with the required properties. Its combined graph is illustrated on Figure \ref{gen:lemma1:fig} below. \end{proof} \begin{figure}[h] \begin{tikzpicture}[xscale=0.4,yscale=0.7] \node[draw,circle,inner sep=1pt,fill] at (3,2) {}; \draw[dashed] (3,2)--(3,1) node[below]{$q_{0}=a$}; \node[draw,circle,inner sep=1pt,fill] at (8,2) {}; \node[draw,circle,inner sep=1pt,fill] at (8,7) {}; \draw[dashed] (8,7)--(8,1) node[below]{$q_{1}$}; \node[draw,circle,inner sep=1pt,fill] at (13,2) {}; \node[draw,circle,inner sep=1pt,fill] at (13,7) {}; \draw[dashed] (13,7)--(13,1) node[below]{$q_{2}$}; \node[draw,circle,inner sep=1pt,fill] at (23,2) {}; \node[draw,circle,inner sep=1pt,fill] at (23,7) {}; \draw[dashed] (23,7)--(23,1) node[below]{$q_{n-1}$}; \node[draw,circle,inner sep=1pt,fill] at (28,7) {}; \draw[dashed] (28,7)--(28,1) node[below]{$q_{n}=b$}; \draw[thick] (3,2) -- (14,2); \draw[dotted, thick] (14,2) -- (15,2); \draw[dotted, thick] (21,2) -- (22,2); \draw[thick] (22,2) -- (23,2); \draw[dashed] (3,2)--(1,2) node[left]{$\displaystyle \frac{a}{n}$}; \draw[thick] (8,7) -- (14,7); \draw[dotted, thick] (14,7) -- (15,7); \draw[dotted, thick] (21,7) -- (22,7); \draw[thick] (22,7) -- (28,7); \draw[dashed] (8,7)--(1,7) node[left]{$\displaystyle \frac{b}{n}$}; \draw[thick] (3,2) -- (8,7); \draw[thick] (8,2) -- (13,7); \draw[thick] (13,2) -- (14,3); \draw[dotted, thick] (14,3) -- (15,4); \draw[thick] (22,6) -- (23,7); \draw[dotted, thick] (21,5) -- (22,6); \draw[thick] (23,2) -- (28,7); \node[below right] at (25.3,4.5) {$P_1$}; \node[below right] at (20.3,4.5) {$P_2$}; \node[below right] at (15.3,4.5) {$P_{n-2}$}; \node[below right] at (10.3,4.5) {$P_{n-1}$}; \node[below right] at (5.3,4.5) {$P_n$}; \end{tikzpicture} \caption{Combined graph of an $n$-system for Lemma \ref{gen:lemma1}.} \label{gen:lemma1:fig} \end{figure} \begin{lemma} \label{gen:lemma2} Let $a,b,c\in\mathbb{R}$ with $0\le a<b<c$. Suppose that $\mathbf{P}^{(1)}\colon[a,b]\to\mathbb{R}^n$ and $\mathbf{P}^{(2)}\colon[b,c]\to\mathbb{R}^n$ are $n$-systems with $\mathbf{P}^{(1)}(b)=\mathbf{P}^{(2)}(b)=(b/n,\dots,b/n)$. Then there is an $n$-system $\mathbf{P}$ on $[a,c]$ which restricts to $\mathbf{P}^{(1)}$ on $[a,b]$ and to $\mathbf{P}^{(2)}$ on $[b,c]$. \end{lemma} \begin{proof} There is a unique map $\mathbf{P}\colon [a,c]\to\mathbb{R}^n$ which restricts to $\mathbf{P}^{(1)}$ on $[a,b]$ and to $\mathbf{P}^{(2)}$ on $[b,c]$. It is continuous but not differentiable at $b$. However, it satisfies the condition (S3) at $q=b$ since all its coordinates are equal at that point. Thus, it is an $n$-system. \end{proof} Let $I$ be any subinterval of $[0,\infty)$ and let $D$ be an arbitrary discrete subset of $I$. By combining Lemmas \ref{gen:lemma1} and \ref{gen:lemma2}, we can construct an $n$-system $\mathbf{P}$ on $I$ which takes the value $\mathbf{P}(q)=(q/n,\dots,q/n)$ at each point $q$ of $D$. Then, by choosing $D$ appropriately, we may ensure that $\sup_{q\in I} \|\mathbf{P}(q)-(q/n,\dots,q/n)\|_\infty$ is finite and arbitrarily small. As we will see, a slightly more general argument shows that these properties of approximation by $n$-systems extend to the following maps, besides the map $\tilde{P}\colon I\to\mathbb{R}^n$ sending each $q\in I$ to $(q/n,\dots,q/n)$. \begin{definition} \label{gen:def2} Let $I$ be a subinterval of $[0,\infty)$ with non-empty interior. A generalized $n$-system on $I$ is a continuous piecewise linear map $\mathbf{P}=(P_1,\dots,P_n)\colon I\to\mathbb{R}^n$ with the following properties. \begin{itemize} \item[(G1)] For each $q\in I$, we have $0\le P_1(q)\le\cdots\le P_n(q)$ and $P_1(q)+\cdots+P_n(q)=q$. \item[(G2)] If $H$ is a non-empty open subinterval of $I$ on which $\mathbf{P}$ is differentiable, then there are integers ${\underline{r}},{\overline{r}}$ with $1\le{\underline{r}}\le{\overline{r}}\le n$ such that $P_{{\underline{r}}},P_{{\underline{r}}+1},\dots,P_{{\overline{r}}}$ coincide on the whole interval $H$ and have slope $1/({\overline{r}}-{\underline{r}}+1)$, while any other component $P_j$ of $\mathbf{P}$ is constant on $H$. \item[(G3)] If $q$ is an interior point of $I$ at which $\mathbf{P}$ is not differentiable, if ${\underline{r}},{\overline{r}},{\underline{s}},\stop$ are the integers for which \begin{equation} \label{gen:def2:eq} P'_j(q^-)=\frac{1}{{\overline{r}}-{\underline{r}}+1} \quad ({\underline{r}}\le j\le{\overline{r}}) \quad\mbox{and}\quad P'_j(q^+)=\frac{1}{\stop-{\underline{s}}+1} \quad ({\underline{s}}\le j\le\stop), \end{equation} and if ${\underline{r}}<\stop$, then we have $P_{{\underline{r}}}(q)=P_{{\underline{r}}+1}(q)=\cdots=P_{\stop}(q)$. \end{itemize} \end{definition} Figure \ref{gen:fig:G3} shows the combined graph of the functions $P_{{\underline{r}}},\dots,P_{\stop}$ over a neighborhood of $q$ under the hypotheses of (G3) when ${\overline{r}}<{\underline{s}}$ and ${\overline{r}}<\stop$. The pattern is slightly different when ${\underline{r}}\ge {\underline{s}}$ (resp.\ when ${\overline{r}}\ge \stop$): in that case, there is no horizontal line segment to the right of $q$ (resp.\ to the left of $q$). However, one cannot have ${\underline{s}}\le {\underline{r}}<\stop\le {\overline{r}}$ and so at least one of the inequalities ${\underline{r}}<{\underline{s}}$ or ${\overline{r}}<\stop$ must hold. \begin{figure}[h] \begin{tikzpicture}[xscale=0.6,yscale=0.3] \node[draw,circle,inner sep=1pt,fill] at (6,4) {}; \draw[dashed] (6,4)--(6,1) node[below]{$q$}; \draw[thick] (3,0.7) -- (6,4) -- (9,6); \draw[thick] (3,4) -- (9,4); \node[left] at (3,0.7) {$P_{{\underline{r}}}=\cdots=P_{{\overline{r}}}$}; \node[left] at (3,4) {$P_{{\overline{r}}+1}=\cdots=P_{\stop}$}; \node[right] at (9,4) {$P_{{\underline{r}}}=\cdots=P_{{\underline{s}}-1}$}; \node[right] at (9,6) {$P_{{\underline{s}}}=\cdots=P_{\stop}$}; \end{tikzpicture} \caption{Illustration for Condition (G3) when ${\overline{r}}<{\underline{s}}$ and ${\overline{r}}<\stop$.} \label{gen:fig:G3} \end{figure} Note also that if \eqref{gen:def2:eq} holds at some point $q$ where $\mathbf{P}$ is not differentiable, and if ${\underline{r}}\ge \stop$, then we must have ${\underline{r}}>\stop$ and $P_{{\underline{r}}}(q)=\cdots=P_{{\overline{r}}}(q) > P_{{\underline{s}}}(q)=\cdots=P_{\stop}(q)$. We leave the verification to the reader. Clearly $n$-systems are also generalized $n$-systems. The next result and its corollary formalize the claims made just above Definition \ref{gen:def2}. \begin{proposition} \label{gen:prop:approx} Let $\tilde{\mathbf{P}}$ be a generalized $n$-system on a subinterval $I$ of $[0,\infty)$ with non-empty interior, and let $D$ be a discrete subset of $I$. Then there is an $n$-system $\mathbf{P}$ on $I$ such that $\mathbf{P}(t)=\tilde{\mathbf{P}}(t)$ for each $t\in D$. \end{proposition} \begin{proof} Let $D_0$ be the set of points of $I$ where $\tilde{\mathbf{P}}$ is not differentiable, including the boundary points of $I$ that lie in $I$ (if any). Since $D_0$ is a discrete subset of $I$, we may assume without loss of generality that $D$ contains $D_0$. We may also assume that $I$ and $D$ have the same infimum and the same supremum. Let $(t_1,t_2)$ be a maximal subinterval of $I\setminus D$. Then, $t_1$ and $t_2$ belong to $D$ and, as $D\subseteq I$, it follows that $[t_1,t_2]$ is a subinterval of $I$. As $\tilde{\mathbf{P}}$ is differentiable on $(t_1,t_2)$, there are integers ${\underline{r}}$ and ${\overline{r}}$ with $1\le {\underline{r}}\le {\overline{r}}\le n$ such that $\tilde{P}_{\underline{r}},\dots,\tilde{P}_{\overline{r}}$ coincide and have slope $1/({\overline{r}}-{\underline{r}}+1)$ on $(t_1,t_2)$ while all other components $\tilde{P}_j$ are equal to a constant $c_j$ on that interval. Put \[ m={\overline{r}}-{\underline{r}}+1 \quad\mbox{and}\quad c=(c_1+\cdots+c_{{\underline{r}}-1})+(c_{{\overline{r}}+1}+\cdots+c_n). \] By Lemma \ref{gen:lemma1}, there exists an $m$-system $(A_1,\dots,A_m)$ on $[t_1,t_2]$ with \[ (A_1(t_i),\dots,A_m(t_i)) = \Big( \frac{t_i}{m},\dots,\frac{t_i}{m} \Big) \quad \text{for $i=1,2$.} \] Since $\tilde{P}_j(q)=(q-c)/m$ for each $j={\underline{r}},\dots,{\overline{r}}$ and each $q\in[t_1,t_2]$, the map $\mathbf{P}\colon[t_1,t_2]\to\mathbb{R}^n$ given by \begin{equation} \label{gen:prop:approx:eq} \mathbf{P}(q) = \Big( c_1, \dots, c_{{\underline{r}}-1}, A_1(q)-\frac{c}{m},\dots,A_m(q)-\frac{c}{m}, c_{{\overline{r}}+1}, \dots, c_n \Big) \quad (t_1\le q\le t_2), \end{equation} coincides with $\tilde{\mathbf{P}}$ at the points $q=t_1,t_2$. As the intervals $[t_1,t_2]$ cover $I$ and have no interior point in common, this in fact defines a continuous piecewise linear map $\mathbf{P}\colon I\to\mathbb{R}^n$ which coincides with $\tilde{\mathbf{P}}$ on the set $D$. To conclude, it remains simply to show that $\mathbf{P}$ satisfies the conditions (S1-S3) of Definition \ref{gen:def1}. For each $q$ in a fixed interval $[t_1,t_2]$, the formula \eqref{gen:prop:approx:eq} shows that the coordinates of $\mathbf{P}(q)$ sum up to $q$. They also form a monotone increasing sequence because $A_1(q)\le\cdots\le A_m(q)$ are separately monotone increasing functions of $q$ in $[t_1,t_2]$, and because the coordinates of $\mathbf{P}(t_i)=\tilde{\mathbf{P}}(t_i)$ form monotone increasing sequences for $i=1,2$. Thus the condition (S1) is fulfilled. The condition (S2) also holds by construction. The same is true for (S3) except possibly at the points $q$ of $D$ which lie in the interior of $I$. The latter points are the common boundary points $q=t_2$ of two maximal subintervals $(t_1,t_2)$ and $(t_2,t_3)$ of $I\setminus D$. Define ${\underline{r}}$, ${\overline{r}}$ as above and denote by ${\underline{s}}$, $\stop$ the corresponding integers for the interval $[t_2,t_3]$. By construction, the function $\mathbf{P}$ satisfies $P'_{\underline{r}}(t_2^-)=P'_\stop(t_2^+)=1$. Suppose that ${\underline{r}}<\stop$. If $t_2\in D_0$, we have $\tilde{P}_{\underline{r}}(t_2)=\cdots=\tilde{P}_\stop(t_2)$ because $\tilde{\mathbf{P}}$ satisfies the condition (G3) of Definition \ref{gen:def2}. If $t_2\notin D_0$, these equalities still hold because then ${\underline{r}}={\underline{s}}$ and ${\overline{r}}=\stop$. As $\mathbf{P}(t_2)=\tilde{\mathbf{P}}(t_2)$, we conclude that $P_{\underline{r}}(t_2)=\cdots=P_\stop(t_2)$ and thus (S3) is satisfied. \end{proof} \begin{corollary} \label{gen:cor} Let $\tilde{\mathbf{P}}\colon I\to\mathbb{R}^n$ be a generalized $n$-system and let $\epsilon>0$. Then there is an $n$-system $\mathbf{P}$ on $I$ such that $\|\tilde{\mathbf{P}}(q)-\mathbf{P}(q)\|_\infty \le \epsilon$ for each $q\in I$. \end{corollary} \begin{proof} Let $D$ be the set of all integer multiples of $\epsilon/2$ in $I$. By Proposition \ref{gen:prop:approx}, there is an $n$-system $\mathbf{P}$ on $I$ such that $\mathbf{P}(t)=\tilde{\mathbf{P}}(t)$ for each $t\in D$. Since the components of $\mathbf{P}$ and of $\tilde{\mathbf{P}}$ have slopes between $0$ and $1$, we find that \[ \|\tilde{\mathbf{P}}(q)-\mathbf{P}(q)\|_\infty \le \|\tilde{\mathbf{P}}(q)-\tilde{\mathbf{P}}(t)\|_\infty + \|\mathbf{P}(q)-\mathbf{P}(t)\|_\infty \le 2|q-t| \] for each $q\in I$ and each $t\in D$. Upon choosing $t$ so that $|q-t|\le \epsilon/2$, we conclude that $\mathbf{P}$ has the required property. \end{proof} Corollary \ref{gen:cor} shows that the set ${\mathcal{S}}_I$ of $n$-systems on a subinterval $I$ of $[0,\infty)$ is dense in the set ${\mathcal{G}}_I$ of generalized $n$-systems on $I$, for the topology of uniform convergence. More precisely, it can be shown that ${\mathcal{G}}_I$ is the completion of ${\mathcal{S}}_I$ for that topology. This justifies working with functions from this set. In particular, the validity of Theorem \ref{gen:thm:nsys} extends to generalized $n$-systems. \section{A family of generalized $n$-systems} \label{sec:fam} Again, we fix an integer $n\ge 2$. Consider the set \[ \Delta^{(n)} = \{ (a_1,\dots,a_n)\in\mathbb{R}^n \,;\, 0<a_1<\cdots<a_n \text{ and } a_1+\cdots+a_n=1 \} \] and its topological closure \[ \bar{\Delta}^{(n)} = \{ (a_1,\dots,a_n)\in\mathbb{R}^n \,;\, 0\le a_1\le\cdots\le a_n \text{ and } a_1+\cdots+a_n=1 \}. \] By condition (G1) from Definition \ref{gen:def2}, we have $q^{-1}\mathbf{P}(q)\in \bar{\Delta}^{(n)}$ for any generalized $n$-system $\mathbf{P}$ and any $q>0$ in its interval of definition. For each $j=1,\dots,n$, we define a map $\psi_j\colon \bar{\Delta}^{(n)} \to \mathbb{R}$ by \[ \psi_j(a_1,\dots,a_n)=a_1+\cdots+a_j. \] The proof of our main result relies on the following basic construction. \begin{proposition} \label{fam:prop:basic} Let $\mathbf{a}=(a_1,\dots,a_n)\in\Delta^{(n)}$. Define \begin{align*} q_i&=a_1+\cdots+a_i+(n-i)a_i \quad (1\le i\le n),\\ q_{n-1+i}&=(i-1)a_i+a_i+\cdots+a_n \quad (1\le i\le n). \end{align*} Then there exists a unique generalized $n$-system $\mathbf{P}=(P_1,\dots,P_n)$ on $[q_1,q_{2n-1}]=[na_1,na_n]$ whose combined graph is as shown on Figure \ref{gen:ex1:fig} below. For each $j=1,\dots,n-1$, it has \[ \inf \psi_j\big(q^{-1}\mathbf{P}(q)\big) = \psi_j(\mathbf{a}) \quad\mbox{and}\quad \sup \psi_j\big(q^{-1}\mathbf{P}(q)\big) = \frac{j}{n} \] where both infimum and supremum are taken over all $q\in[na_1,na_n]$. \end{proposition} In Figure \ref{gen:ex1:fig}, the number $1/m$ next to each slanted line segment indicates its slope and thus the number $m$ of components of $\mathbf{P}$ whose graph coincide with this line segment over the corresponding interval $[q_i,q_{i+1}]$. \begin{figure}[h] \begin{tikzpicture}[scale=0.35] \node[draw,circle,inner sep=1pt,fill] at (4,3) {}; \draw[dashed] (4,3)--(4,2) node[below]{$q_{1}$}; \node[draw,circle,inner sep=1pt,fill] at (9,5) {}; \draw[dashed] (9,5)--(9,2) node[below]{$q_{2}$}; \node[draw,circle,inner sep=1pt,fill] at (13,7) {}; \draw[dashed] (13,7)--(13,2) node[below]{$q_{3}$}; \node[draw,circle,inner sep=1pt,fill] at (18,11) {}; \draw[dashed] (18,11)--(18,2); \node[below] at (17.5,2) {$q_{n-1}$}; \node[draw,circle,inner sep=1pt,fill] at (20,15) {}; \node[draw,circle,inner sep=1pt,fill] at (20,3) {}; \draw[dashed] (20,15)--(20,2); \node[below] at (19.75,2) {$q_{n}$}; \node[draw,circle,inner sep=1pt,fill] at (21,5) {}; \draw[dashed] (21,5)--(21,2); \node[below] at (21.6,2) {$q_{n+1}$}; \node[draw,circle,inner sep=1pt,fill] at (23,7) {}; \draw[dashed] (23,7)--(23,2); \node[below] at (23.9,2) {$q_{n+2}$}; \node[draw,circle,inner sep=1pt,fill] at (30,11) {}; \draw[dashed] (30,11)--(30,2) node[below]{$q_{2n-2}$}; \node[draw,circle,inner sep=1pt,fill] at (40,15) {}; \draw[dashed] (40,15)--(40,2) node[below]{$q_{2n-1}$}; \draw[dashed] (4,3)--(2,3) node[left]{$a_{1}$}; \draw[dashed] (9,5)--(2,5) node[left]{$a_{2}$}; \draw[dashed] (13,7)--(2,7) node[left]{$a_{3}$}; \node[left] at (1.5,9) {$\vdots$}; \draw[dashed] (18,11)--(2,11) node[left]{$a_{n-1}$}; \draw[dashed] (20,15)--(2,15) node[left]{$a_{n}$}; \draw[thick] (4,3) -- (20,3); \draw[thick] (9,5) -- (21,5); \draw[thick] (13,7) -- (23,7); \draw[thick] (18,11) -- (30,11); \draw[thick] (20,15) -- (40,15); \draw[thick] (20,3) -- (21,5) -- (23,7) -- (24,23/3); \draw[dotted, thick] (24,23/3) -- (25,25/3); \draw[dotted, thick] (27,19/2) -- (28,10); \draw[thick] (28,10) -- (30,11) -- (40,15); \draw[thick] (4,3) -- (9,5) -- (13,7) -- (14,23/3); \node[font=\footnotesize] at (4.5,4.2) {$\frac{1}{n-1}$}; \node[font=\footnotesize] at (9.3,6.2) {$\frac{1}{n-2}$}; \node[font=\footnotesize] at (13.2,8.3) {$\frac{1}{n-3}$}; \node[font=\footnotesize] at (15.9,10) {$\frac{1}{2}$}; \node[font=\footnotesize] at (18.2,13) {$1$}; \draw[dotted, thick] (14,23/3) -- (15,25/3); \draw[dotted, thick] (16.3,9.3) -- (17,10); \draw[thick] (17,10) -- (18,11) -- (20,15) -- (40,15); \node[font=\footnotesize] at (20.6,3.2) {$1$}; \node[font=\footnotesize] at (22.3,5.3) {$\frac{1}{2}$}; \node[font=\footnotesize] at (24.6,7.2) {$\frac{1}{3}$}; \node[font=\footnotesize] at (28,9) {$\frac{1}{n-2}$}; \node[font=\footnotesize] at (36,12.5) {$\frac{1}{n-1}$}; \end{tikzpicture} \caption{The combined graph of a generalized $n$-system.} \label{gen:ex1:fig} \end{figure} The existence and uniqueness of $\mathbf{P}$ is easily seen. In order to compute the required infimum and supremum, we use the following result. \begin{lemma} \label{fam:lemmaM} Let $0<a<b$, let $M\colon [a,b]\to\mathbb{R}$ be a continuous piecewise linear function on $[a,b]$, let $q_1=a<q_2<\cdots<q_t=b$ be the points where $M$ is not differentiable (including $a$ and $b$), and let $\rho_i$ denote the constant value of the derivative of $M$ on $(q_i,q_{i+1})$ for $i=1,\dots,t-1$. \begin{itemize} \item[1)] If $M(a)/a>\rho_1>\cdots>\rho_{t-1}$, then $M(q)/q$ is strictly decreasing on $[a,b]$, and bounded below by $\rho_{t-1}$. \item[2)] If $\rho_1>\cdots>\rho_{t-1}>M(b)/b$, then $M(q)/q$ is strictly increasing on $[a,b]$, and bounded above by $\rho_1$. \end{itemize} \end{lemma} \begin{proof} It suffices to prove this for $t=2$ because the general case then follows by induction on $t$. Assuming that $t=2$, we can write \[ M(q)=M(a)+\rho_1(q-a)=M(b)+\rho_1(q-b) \quad (a\le q\le b). \] Under the hypothesis 1), we also have $M(a)>\rho_1a$ and so $M(q)/q=\rho_1+(M(a)-\rho_1a)/q$ is a strictly decreasing function of $q$ on $[a,b]$, bounded below by $\rho_1$. Similarly, under the hypothesis 2), we have $M(b)<\rho_1b$, then $M(q)/q=\rho_1+(M(b)-\rho_1b)/q$ is strictly increasing on $[a,b]$ and bounded above by $\rho_1$. \end{proof} \begin{proof}[Proof of Proposition \ref{fam:prop:basic}] As noted above, it suffices to prove the second assertion. To this end, fix an index $j$ with $1\le j\le n-1$ and define $M_j=P_1+\dots+P_j$. Then $M_j$ is a continuous piecewise linear map on $[q_1,q_{2n-1}]$ with slope $(j-i)/(n-i)$ on $[q_i,q_{i+1}]$ for $i=1,\dots,j-1$, slope $0$ on $[q_j,q_n]$, slope $1$ on $[q_n,q_{n+j}]$ and slope $j/i$ on $[q_{n+i-1},q_{n+i}]$ for $i=j+1,\dots,n-1$. Since \[ \frac{M_j(q_1)}{q_1} =\frac{j}{n}>\frac{j-1}{n-1}>\cdots>\frac{1}{n-j+1}>0 \] and \[ 1>\frac{j}{j+1}>\cdots>\frac{j}{n-1}>\frac{j}{n} =\frac{M_j(q_{2n-1})}{q_{2n-1}}, \] we conclude from Lemma \ref{fam:lemmaM} that the ratio $M_j(q)/q$ is strictly decreasing on $[q_1,q_n]$ and strictly increasing on $[q_n,q_{2n-1}]$. This means that \[ \inf\frac{M_j(q)}{q} =\frac{M_j(q_n)}{q_n} =\frac{a_1+\cdots+a_j}{a_1+\cdots+a_n} =\psi_j(\mathbf{a}) \quad\mbox{and}\quad \sup \frac{M_j(q)}{q} =\frac{j}{n}. \qedhere \] \end{proof} The last result of this section combines the above proposition with the following observation. \begin{lemma} \label{fam:lemma:rescaling} If $\mathbf{P}\colon [a,b]\to\mathbb{R}^n$ is a generalized $n$-system with $b>a>0$, then, for any $d>c>0$ with $d/c=b/a$, the map $\tilde{\mathbf{P}} \colon [c,d]\to\mathbb{R}^n$ given by \[ \tilde{\mathbf{P}}(q) = \frac{c}{a} \mathbf{P}\Big(\frac{a}{c}q\Big) \quad \text{for each $q\in[c,d]$} \] is also a generalized $n$-system and we have both \[ \inf_{q\in[a,b]}\psi_j\big(q^{-1}\mathbf{P}(q)\big) = \inf_{q\in[c,d]}\psi_j\big(q^{-1}\tilde{\mathbf{P}}(q)\big) \quad\mbox{and}\quad \sup_{q\in[a,b]}\psi_j\big(q^{-1}\mathbf{P}(q)\big) = \sup_{q\in[c,d]}\psi_j\big(q^{-1}\tilde{\mathbf{P}}(q)\big). \] \end{lemma} The combined graph of this new map $\tilde{\mathbf{P}}$ is the image of the combined graph of $\mathbf{P}$ by uniform scaling of ratio $c/a$. Upon noting furthermore that Lemma \ref{gen:lemma2} extends to generalized $n$-systems, we conclude with the following result. \begin{proposition} \label{fam:prop:pasting} Let $E$ be an arbitrary non-empty subset of $\bar{\Delta}^{(n)}$. There exists a generalized $n$-system $\mathbf{P}$ on $[1,\infty)$ which, for each $j=1,\dots,n$, satisfies \[ \liminf_{q\to\infty} \psi_j\big(q^{-1}\mathbf{P}(q)\big) = \inf \psi_j(E) \quad\mbox{and}\quad \limsup_{q\to\infty} \psi_j\big(q^{-1}\mathbf{P}(q)\big) = \frac{j}{n}. \] \end{proposition} \begin{proof} If $E$ consists of the single point $(1/n,\dots,1/n)$, we simply take $\mathbf{P}(q)=(q/n,\dots,q/n)$ for each $q\ge 1$. Otherwise, we choose a sequence $(\mathbf{a}^{(i)})_{i\ge 1}$ of points of $\Delta^{(n)}$ whose set of accumulation points is the topological closure $\bar{E}$ of $E$. For each $i\ge 1$, we consider the generalized $n$-system attached to $\mathbf{a}^{(i)}$ by Proposition \ref{fam:prop:basic}, and, starting with $q_1=1$, we use Lemma \ref{fam:lemma:rescaling} recursively to transform it into a generalized $n$-system $\mathbf{P}^{(i)}$ on an interval of the form $[q_i,q_{i+1}]$. By construction, we have \begin{equation} \label{fam:prop:pasting:eq} \inf_{q\in[q_i,q_{i+1}]} \psi_j\big(q^{-1}\mathbf{P}^{(i)}(q)\big) = \psi_j(\mathbf{a}^{(i)}) \quad\mbox{and}\quad \sup_{q\in[q_i,q_{i+1}]} \psi_j\big(q^{-1}\mathbf{P}^{(i)}(q)\big) = \frac{j}{n} \end{equation} for $j=1,\dots,n$. We also note that \[ \mathbf{P}^{(i-1)}(q_i)=\mathbf{P}^{(i)}(q_i)=(q_i/n,\dots,q_i/n) \quad (i\ge 2) \] and that $\limsup q_{i+1}/q_i >1$ since $E$ contains at least one point $(a_1,\dots,a_n)$ with $a_n>a_1$. Thus $\lim q_i=\infty$ and so there exists a unique generalized $n$-system $\mathbf{P}$ on $[1,\infty)$ which restricts to $\mathbf{P}^{(i)}$ on $[q_i,q_{i+1}]$ for each $i\ge 1$. In view of \eqref{fam:prop:pasting:eq}, we obtain \[ \liminf_{q\to\infty} \psi_j\big(q^{-1}\mathbf{P}(q)\big) = \liminf_{q\to\infty} \psi_j(\mathbf{a}^{(i)}) = \inf \psi_j(E) \quad\mbox{and}\quad \limsup_{q\to\infty} \psi_j\big(q^{-1}\mathbf{P}(q)\big)= \frac{j}{n} \] for each $j=1,\dots,n$, as requested. \end{proof} \section{Proof of the main result} \label{sec:proof} The next result is the last piece that we need in order to prove Theorem \ref{results:thm:main}. \begin{proposition} \label{proof:prop1} Let $n\in\mathbb{N}^*$. Suppose that ${\ubar{\psi}}_1,\dots,{\ubar{\psi}}_n \in [0,1]$ satisfy ${\ubar{\psi}}_n\le n/(n+1)$, \begin{equation} \label{proof:prop1:eq} \frac{{\ubar{\psi}}_j}{j} \le \frac{{\ubar{\psi}}_{j+1}}{j+1} \quad\mbox{and}\quad \frac{1-{\ubar{\psi}}_j}{n+1-j} \le \frac{1-{\ubar{\psi}}_{j+1}}{n-j} \quad (1\le j\le n-1). \end{equation} Then, there exists a finite non-empty subset $E$ of $\bar{\Delta}^{(n+1)}$ such that \[ {\ubar{\psi}}_j=\min \psi_j(E) \quad (1\le j\le n). \] \end{proposition} \begin{proof} If $n=1$, we have ${\ubar{\psi}}_1\le 1/2$ and we simply take $E=\{({\ubar{\psi}}_1,1-{\ubar{\psi}}_1)\}$. Suppose from now on that $n\ge 2$ and let $k\in\{1,\dots,n-1\}$. We claim that there exists $\mathbf{a}\in\bar{\Delta}^{(n+1)}$ such that $\psi_j(\mathbf{a})\ge {\ubar{\psi}}_j$ for $j=1,\dots,n$, with equality for $j=k$ and $j=k+1$. To show this, set $c={\ubar{\psi}}_k/k$ and $d=(1-{\ubar{\psi}}_{k+1})/(n-k)$, and consider the point \[ \mathbf{a} = \big( \overbrace{c,\dots,c}^{\text{$k$ times}},\ {\ubar{\psi}}_{k+1}-{\ubar{\psi}}_k, \overbrace{d,\dots,d}^{\text{$n-k$ times}} \big) \in \mathbb{R}^{n+1}. \] By hypothesis, we have $c\ge 0$ and the inequalities \eqref{proof:prop1:eq} for $j=k$ yield $c \le {\ubar{\psi}}_{k+1}-{\ubar{\psi}}_k \le d$. As the sum of the coordinates of $\mathbf{a}$ is $1$, this means that $\mathbf{a}\in\bar{\Delta}^{(n+1)}$. We also note from \eqref{proof:prop1:eq} that the ratios ${\ubar{\psi}}_j/j$ and $(1-{\ubar{\psi}}_j)/(n+1-j)$ are both monotone increasing functions of $j$ for $j\in\{1,\dots,n\}$. Thus, for $j=1,\dots,k$, we have ${\ubar{\psi}}_j/j\le c$ and so we find $\psi_j(\mathbf{a})=jc \ge {\ubar{\psi}}_j$, with equality if $j=k$. Similarly, for $j=k+1,\dots,n$, we have $d\le (1-{\ubar{\psi}}_j)/(n+1-j)$ and so we obtain $\psi_j(\mathbf{a})=1-(n+1-j)d \ge {\ubar{\psi}}_j$, with equality if $j=k+1$. This proves our claim. By varying $k$, this produces a finite set of points with the required property. \end{proof} In general, we cannot expect that the set $E$ consists of a single element because each $\mathbf{a}=(a_1,\dots,a_{n+1})$ in $\bar{\Delta}^{(n+1)}$ satisfies \[ \psi_{j-1}(\mathbf{a})+\psi_{j+1}(\mathbf{a})-2\psi_j(\mathbf{a}) = a_{j+1}-a_j \ge 0 \quad (2\le j\le n), \] and so, for example, there is no point $\mathbf{a}$ in $\bar{\Delta}^{(4)}$ with $\psi_1(\mathbf{a})=0$, $\psi_2(\mathbf{a})=1/3$ and $\psi_3(\mathbf{a})=1/2$, although the numbers ${\ubar{\psi}}_1=0$, ${\ubar{\psi}}_2=1/3$ and ${\ubar{\psi}}_3=1/2$ satisfy the conditions \eqref{proof:prop1:eq} for $n=3$. \begin{proof}[Proof of Theorem \ref{results:thm:main}] Suppose that $\omega_0,\dots,\omega_{n-1}\in[0,\infty]$ satisfy the conditions of the theorem. Then the numbers \[ {\ubar{\psi}}_j = \frac{1}{\omega_{n-j}+1} \in [0,1] \quad (1\le j\le n), \] fulfill the hypotheses of the above proposition. Let $E$ be a corresponding subset of $\bar{\Delta}^{(n+1)}$, as in that proposition, and let $\mathbf{P}$ be a generalized $(n+1)$-system on $[1,\infty)$ as constructed by Proposition \ref{fam:prop:pasting} for this choice of $E$. For $j=1,\dots,n$, we have \[ \liminf_{q\to\infty} \psi_j\big(q^{-1}\mathbf{P}(q)\big) = \inf \psi_j(E) = {\ubar{\psi}}_j \quad\mbox{and}\quad \limsup_{q\to\infty} \psi_j\big(q^{-1}\mathbf{P}(q)\big) = \frac{j}{n+1}. \] Corollary \ref{gen:cor} and Theorem \ref{gen:thm:nsys} provide a point $\mathbf{u}\in\mathbb{R}^{n+1}\setminus\{0\}$ for which the difference $\mathbf{P}-\mathbf{L}_\mathbf{u}$ is bounded on $[1,\infty)$. By the above, and by definition of the quantities ${\ubar{\psi}}_j(\mathbf{u})$ and ${\bar{\psi}}_j(\mathbf{u})$ in Section \ref{sec:link}, this point satisfies \[ {\ubar{\psi}}_j(\mathbf{u})={\ubar{\psi}}_j \quad\mbox{and}\quad {\bar{\psi}}_j(\mathbf{u})=\frac{j}{n+1} \quad (1\le j\le n). \] By Proposition \ref{link:prop:omega}, this means that, as requested, we have, for $j=0,\dots,n-1$, \[ \omega_j(\mathbf{u}) =\frac{1}{{\ubar{\psi}}_{n-j}(\mathbf{u})}-1 =\omega_j \quad\mbox{and}\quad \hat{\omega}_j(\mathbf{u}) =\frac{1}{{\bar{\psi}}_{n-j}(\mathbf{u})}-1 =\frac{j+1}{n-j}. \] Since $\hat{\omega}_{n-1}(\mathbf{u})=n<\infty$, the coordinates of $\mathbf{u}$ must be linearly independent over $\mathbb{Q}$. \end{proof} \medskip As a complement, note that each point $\mathbf{a}=(a_1,\dots,a_{n+1})$ in $\bar{\Delta}^{(n+1)}$ satisfies \begin{equation} \label{proof:complement:eq} 0 \le \frac{\psi_j(\mathbf{a})}{j} \le \frac{\psi_{j+1}(\mathbf{a})}{j+1} \quad\mbox{and}\quad 0 \le \frac{1-\psi_j(\mathbf{a})}{n+1-j} \le \frac{1-\psi_{j+1}(\mathbf{a})}{n-j} \quad (1\le j\le n-1). \end{equation} Based on this, Theorem \ref{results:thm:SL} can be given the following short proof. Let $\mathbf{u}\in\mathbb{R}^{n+1}\setminus\{0\}$. By Theorem \ref{gen:thm:nsys}, there exists $q_0>0$ and an $(n+1)$-system $\mathbf{P}$ on $[q_0,\infty)$ such that $\mathbf{P}-\mathbf{L}_\mathbf{u}$ is bounded on $[q_0,\infty)$. For such a choice of $\mathbf{P}$, we have \[ {\ubar{\psi}}_j(\mathbf{u}) = \liminf_{q\to\infty} \psi_j\big(q^{-1}\mathbf{P}(q)\big) \quad (1\le j\le n). \] Since $q^{-1}\mathbf{P}(q)$ belongs to $\bar{\Delta}^{(n+1)}$ for each $q\ge q_0$, the inequalities \eqref{proof:complement:eq} apply to these points. They imply that \[ 0 \le \frac{{\ubar{\psi}}_j(\mathbf{u})}{j} \le \frac{{\ubar{\psi}}_{j+1}(\mathbf{u})}{j+1} \quad\mbox{and}\quad 0 \le \frac{1-{\ubar{\psi}}_j(\mathbf{u})}{n+1-j} \le \frac{1-{\ubar{\psi}}_{j+1}(\mathbf{u})}{n-j} \quad (1\le j\le n-1), \] and the conclusion follows using Proposition \ref{link:prop:omega} to translate these estimates in terms of $\omega_0(\mathbf{u}),\dots,\omega_{n-1}(\mathbf{u})$.
\section{Introduction} According to quark models (QM's) \cite{Isgur:1978xj,Capstick:1986bm,Bijker:1994yr,Ferraris:1995ui,Glozman:1995fu,Loring:2001kv}, baryons can be described as the bound states of three constituent quarks. These are effective degrees of freedom that mimic the three valence quarks inside baryons, with a sea of gluons and $q \bar q$ sea pairs. The light baryons can then be ordered according to the approximate SU$_{\mbox{f}}$(3) symmetry into the multiplets $[1]_A \oplus [8]_M \oplus [8]_M \oplus [10]_S$. QM's explain quite well several properties of baryons, such as the strong decays and the magnetic moments. Nevertheless, they predict a larger number of states than the experimentally observed ones (the missing resonances problem) and states with certain quantum numbers appear in the spectrum at excitation energies much lower than predicted \cite{Nakamura:2010zzi}. The problem of the missing resonances \cite{Nakamura:2010zzi,Capstick:1992uc,Capstick:1992th} has motivated the realization of several experiments, such as CB-ELSA \cite{Crede:2003ax}, TAPS \cite{Krusche:1995nv}, GRAAL \cite{Renard:2000iv}, SAPHIR \cite{Tran:1998qw} and CLAS \cite{Dugger:2002ft}, which only provided a few weak indications about some states. Indeed, even if several experiments have been dedicated to the search of missing states, just a small number of new resonances has been included into the PDG \cite{Nakamura:2010zzi}. There are two possible explanations to the puzzle of the missing resonances: 1) There may be resonances very weakly coupled to the single pion, but with higher probabilities of decaying into two or more pions or into other mesons \cite{Nakamura:2010zzi,Capstick:1992uc,Capstick:1992th}. The detection of such states is further complicated by the problem of the separation of the experimental data from the background and by the expansion of the differential cross section into many partial waves; 2) Alternately, it is possible to consider models that are characterized by a smaller number of effective degrees of freedom with respect to the three quarks QM's and to assume that the majority of the missing states, not yet experimentally observed, simply may not exist. This is the case of quark-diquark models \cite{Ida:1966ev,Anselmino:1992vg,Ferretti:2011zz,DeSanctis:2011zz, Galata:2012xt,Jaffe:2004ph,Wilczek:2004im,Selem:2006nd,Santopinto:2004hw,Cloet:2009,Forkel:2008un,Anisovich:2010wx}, where two quarks are strongly correlated and thus the state space is heavily reduced. In quark-diquark models, the effective degrees of freedom of diquarks are introduced to describe baryons as bound states of a constituent diquark and quark \cite{Ida:1966ev}. The notion of diquark dates back to 1964, when its possibility was mentioned by Gell-Mann \cite{GellMann:1964nj} in his original paper on quarks. Since then, many papers have been written on this topic (for a review see Ref. \cite{Anselmino:1992vg}) and, more recently, the diquark concept has been applied to various calculations \cite{Jakob:1997,Bloch:1999ke,Brodsky:2002,Ma,Oettel:2002wf,Gamberg:2003,Jaffe:2003sg,Wilczek:2004im,Maris:2004,Jaffe:2004ph,Selem:2006nd, DeGrand:2007vu,BacchettaRadici,Forkel:2008un,Cloet:2009,Anisovich:2010wx,Santopinto:2004hw,Ferretti:2011zz,DeSanctis:2011zz,Galata:2012xt}. Important phenomenological indications for diquark-like correlations have been collected \cite{Jaffe:2004ph,Selem:2006nd,Close:1988br,Neubert} and indications for diquark confinement have also been provided \cite{Bender:1996bb}. This makes plausibly enough to make diquarks a part of the baryon's wave function. In Ref. \cite{Santopinto:2004hw}, one of us developed an nonrelativistic interacting quark-diquark model, i.e. a potential model based on the effective degrees of freedom of a constituent quark and diquark. In Ref. \cite{Ferretti:2011zz}, it was "relativized" and reformulated within the point form formalism \cite{Klink:1998zz}. In Ref. \cite{DeSanctis:2011zz}, we used the wave functions of Ref. \cite{Ferretti:2011zz} to compute the nucleon electromagnetic form factors. Here, we intend to improve the "relativized" model \cite{Ferretti:2011zz,DeSanctis:2011zz} and compute the non strange baryon spectrum within point form dynamics. Even if our previous results for the non strange baryon spectrum \cite{Ferretti:2011zz} were in general quite good, here we intend to show how the introduction of a spin-isospin transition interaction, inducing the mixing between quark-scalar diquark and quark-axial-vector diquark states in the nucleon wave function, can further improve them, as already suggested in Ref. \cite{Santopinto:2004hw}. Scalar and axial-vector diquarks are two correlated quarks in $S$ wave with spin 0 or 1, respectively \cite{Wilczek:2004im,Jaffe:2004ph}. In a following paper, we will use the new wave functions, obtained by solving the eigenvalue problem of the mass operator of the present model, to compute the nucleon electromagnetic form factors and the elicity amplitudes. \section{The Mass operator} \label{The Model} We consider a quark-diquark system, where $\vec{r}$ is the relative coordinate between the two constituents and $\vec{q}$ is the conjugate momentum to $\vec{r}$. We propose a relativistic quark-diquark model, based on the following baryon rest frame mass operator \begin{equation} \begin{array}{rcl} M & = & E_0 + \sqrt{\vec q\hspace{0.08cm}^2 + m_1^2} + \sqrt{\vec q\hspace{0.08cm}^2 + m_2^2} + M_{\mbox{dir}}(r) \\ & + & M_{\mbox{cont}}(q,r) + M_{\mbox{ex}}(r) + M_{\mbox{tr}}(r)~, \end{array} \label{eqn:H0} \end{equation} where $E_0$ is a constant, $M_{\mbox{dir}}(r)$ and $M_{\mbox{ex}}(r)$ respectively the direct and the exchange diquark-quark interaction, $m_1$ and $m_2$ stand for diquark and quark masses, where $m_1$ is either $m_S$ or $m_{AV}$ according if the part of the mass operator diagonal in the diquark spin [i.e. the whole mass operator of Eq. (\ref{eqn:H0}) without the interaction $M_{\mbox{tr}}(r)$] acts on a scalar or an axial-vector diquark \cite{Jaffe:2004ph,Lichtenberg:1979de,deCastro:1993sr,Schafer:1993ra,Cahill:1995ka,Lichtenberg:1996fi,Burden:1996nh,Maris,Orginos:2005vr,Wilczek:2004im,Flambaum:2005kc,Eichmann:2008ef,Babich:2007ah}, $M_{\mbox{cont}}(q,r)$ is a contact interaction and $M_{\mbox{tr}}(r)$ is a spin-isospin transition interaction. The direct term is a Coulomb-like interaction with a cut off plus a linear confinement term \begin{equation} \label{eq:Vdir} M_{\mbox{dir}}(r)=-\frac{\tau}{r} \left(1 - e^{-\mu r}\right)+ \beta r ~~. \end{equation} One needs also an exchange interaction \cite {Lichtenberg:1981pp,Santopinto:2004hw}, since this is the crucial ingredient of a quark-diquark description of baryons. We have \begin{gather} M_{\mbox{ex}}(r) = (-1)^{l + 1} e^{-\sigma r} \left [ A_S \mbox{ } \vec s_1 \cdot \vec s_2 + A_I \mbox{ } \vec t_1 \cdot \vec t_2 \right . \notag \\ +\left . A_{SI} \mbox{ } \vec s_1 \cdot \vec s_2 \mbox{ } \vec t_1 \cdot \vec t_2 \right ] ~~, \label{eqn:VexchS1} \end{gather} where $\vec{s}$ and $\vec{t}$ are the spin and the isospin operators. Moreover, we consider a contact interaction similar to that introduced by Godfrey and Isgur \cite{Godfrey:1985xj} \begin{equation} \label{eqn:Vcont(r)} \begin{array}{rcl} M_{\mbox{cont}} & = & \left(\frac{m_1 m_2}{E_1 E_2}\right)^{1/2} \frac{\eta^3 D}{\pi^{3/2}} e^{-\eta^2 r^2} \mbox{ } \delta_{L,0} \delta_{s_1,1} \left(\frac{m_1 m_2}{E_1 E_2}\right)^{1/2} \mbox{ }, \end{array} \end{equation} where $E_i = \sqrt{\vec q\hspace{0.08cm}^2 + m_i^2}$ ($i$ = 1, 2), $\eta$ and $D$ are parameters of the model. Finally we consider a spin-isospin transition interaction, $M_{\mbox{tr}}(r)$, in order to mix quark-scalar diquark and quark-axial-vector diquark states. $M_{\mbox{tr}}(r)$ is chosen as \begin{equation} \label{eqn:Vtr(r)} M_{\mbox{tr}}(r) = V_0 \mbox{ } e^{-\frac{1}{2} \nu^2 r^2} (\vec s_2 \cdot \vec S) (\vec t_2 \cdot \vec T) \mbox{ }, \end{equation} where $V_0$ and $\nu$ are free parameters. The matrix elements of the spin transition operator, $\vec S$, are defined as: \begin{subequations} \begin{equation} \left\langle \right. s_1', m_{s_1}' \left. \right| S_\mu^{[1]} \left| s_1, m_{s_1} \right\rangle \neq 0 \mbox{ for } s_1' \neq s_1 \mbox{ }, \end{equation} where \begin{equation} \left\langle 1 \right\| S_1 \left\| 0 \right\rangle = 1 \mbox{ }, \end{equation} \begin{equation} \left\langle 0 \right\| S_1 \left\| 1 \right\rangle = -1 \end{equation} \end{subequations} and the same holds for those of the isospin transition operator, $\vec T$. Thus one has: \begin{equation} \begin{array}{rcl} \left\langle \Phi' \right| M_{\mbox{tr}} \left| \Phi \right\rangle & = & \frac{1}{4} V_0 \delta_{s_1',s_1\pm1} \delta_{S\frac{1}{2}} \delta_{t_1',t_1\pm1} \delta_{T\frac{1}{2}} \\ & \times & \left\langle \Phi'(\vec r) \right| e^{-\frac{1}{2} \nu^2 r^2} \left| \Phi(\vec r) \right\rangle \mbox{ }, \end{array} \end{equation} where $\Phi(\vec r)$ stands for the spatial wave function of the generic state, $\left| \Phi \right\rangle$. The mass formula of the previous version of the relativistic quark-diquark model \cite{Ferretti:2011zz} is \begin{widetext} \begin{equation} \label{eqn:RMF} \begin{array}{rcl} M & = & E_0 + \sqrt{\vec q\hspace{0.08cm}^2 + m_1^2} + \sqrt{\vec q\hspace{0.08cm}^2 + m_2^2} - \frac{\tau}{r} \left(1 - e^{-\mu r}\right) + \beta r + \left(\frac{m_1 m_2}{E_1 E_2}\right)^{1/2+\epsilon} \frac{\eta^3 D}{\pi^{3/2}} e^{-\eta^2 r^2} \mbox{ } \delta_{L,0} \delta_{s_1,1} \left(\frac{m_1 m_2}{E_1 E_2}\right)^{1/2+\epsilon} \\ & + & (-1)^{l + 1} e^{-\sigma r} \left [ A_S \mbox{ } \vec s_1 \cdot \vec s_2 + A_I \mbox{ } \vec t_1 \cdot \vec t_2 + A_{SI} \mbox{ } \vec s_1 \cdot \vec s_2 \mbox{ } \vec t_1 \cdot \vec t_2 \right ] ~~. \end{array} \end{equation} \end{widetext} The main difference between the mass operator of Eq. (\ref{eqn:H0}) and that of Eq. (\ref{eqn:RMF}) \cite{Ferretti:2011zz} is the presence of the spin-isospin transition interaction $M_{\mbox{tr}}$ in Eq. (\ref{eqn:H0}). $M_{\mbox{tr}}(r)$ is introduced to improve the description of the electromagnetic elastic form factors of the nucleon \cite{DeSanctis:2011zz,DeSanctis:TBP}. Indeed, $M_{\mbox{tr}}(r)$ makes it possible to have a nucleon wave function with a quark-axial-vector diquark component in addition to the quark-scalar diquark one. At the same time, $M_{\mbox{tr}}(r)$ significantly improves the description of the non strange baryon spectrum \cite{Ferretti:2011zz} (see Fig. \ref{fig:Spectrum3e4}). One can also notice that the values of the model parameters change significantly from those of Ref. \cite{DeSanctis:2011zz,Ferretti:2011zz} after the introduction of the interaction (\ref{eqn:Vtr(r)}) into the mass formula. In particular, one can see that the masses of the two constituents (the quark and the diquark) are now smaller than before, which is good in a relativistic QM, and the mass difference between the scalar and the axial-vector diquark is smaller too (it goes from 350 MeV to 210 MeV). The same happens for the string tension, that goes from 2.15 $\mbox{fm}^{-2}$ to 1.57 $\mbox{fm}^{-2}$. It is worth noting that the number of model parameters increases only by one, since there are two new parameters, $V_0$ and $\nu$ [see Eq. (\ref{eqn:Vtr(r)})], while the parameter $\epsilon$ of the contact interaction [see Eqs. (\ref{eqn:Vcont(r)}) and (\ref{eqn:RMF})] has been removed. \begin{table}[h] \begin{center} \begin{tabular}{llllll} \hline \hline \\ $m_q$ & $=140$ MeV & $~m_{S}$ & $=150$ MeV & $~m_{AV}$ & $=360$ MeV \\ $~\tau$ & $=1.23$ & $~\mu$ & $=125~\mbox{fm}^{-1}$ & $~\beta$ & $=1.57~\mbox{fm}^{-2}$ \\ $A_S$ & $=125$ MeV & $A_I$ & $=85$ MeV & $A_{SI}$ & $=350$ MeV \\ $~\sigma$ & $=0.60~\mbox{fm}^{-1}$ & $~E_0$ & $=826$ MeV & $D$ & $=2.00$ $\mbox{fm}^2$ \\ $~\eta$ & $=10.0~\mbox{fm}^{-1}$ & $~V_0$ & $=1450$ MeV & $\nu$ & $=0.35$ $\mbox{fm}^{-1}$ \\ \\ \hline \hline \end{tabular} \end{center} \caption{Resulting values for the model parameters.} \label{tab:ResultingParameters} \end{table} Finally, it has to be noted that in the present work all the calculations are performed without any perturbative approximation, as in Ref. \cite{Ferretti:2011zz}. The eigenfunctions of the mass operator of Eq. (\ref{eqn:H0}) can be thought as eigenstates of the mass operator with interaction in a Bakamjian-Thomas construction \cite{BT}. The interaction is introduced adding an interaction term to the free mass operator $M_0 = \sqrt{\vec q\hspace{0.08cm}^2 + m_1^2} + \sqrt{\vec q\hspace{0.08cm}^2 + m_2^2}$, in such a way that the interaction commutes with the non interacting Lorenz generators and with the non interacting four velocity \cite{KP}. The dynamics is given by a point form Bakamjian-Thomas construction. Point form means that the Lorentz group is kinematic. Furthermore, since we are doing a point form Bakamjian-Thomas construction, here $P = M V_0$ where $V_0$ is the noninteracting four-velocity (whose eigenvalue is $v$). The general quark-diquark state, defined on the product space $H_1 \otimes H_2$ of the one-particle spin $s_1$ (0 or 1) and spin $s_2$ (1/2) positive energy representations $H_1=L^2(R^3)\otimes S_1^{0}$ or $H_1=L^2(R^3)\otimes S_1^{1}$ and $H_2=L^2(R^3)\otimes S_2^{1/2}$ of the Poincar\'e Group, can be written as \cite{Ferretti:2011zz} \begin{equation} \left| p_1, p_2, \lambda_1, \lambda_2 \right\rangle \mbox{ }, \end{equation} where $p_1$ and $p_2$ are the four-momenta of the diquark and the quark, respectively, while $\lambda_1$ and $\lambda_2$ are, respectively, the $z$-projections of their spins. We introduce the velocity states as \cite{Klink:1998zz,Ferretti:2011zz} \begin{equation} \label{eqn:velocity-states} \vert v,\vec{k}_1,\lambda_1,\vec{k}_2,\lambda_2\rangle = U_{B(v)}\vert k_1,s_1,\lambda_1, k_2,s_2,\lambda_2 \rangle_{0} \mbox{ }, \end{equation} where the suffix $0$ means that the diquark and the quark three-momenta $\vec {k}_1$ and $\vec{k}_2$, called internal momenta, satisfy: \begin{equation} \vec {k}_1 + \vec{k}_2=0 \mbox{ }. \end{equation} Following the standard rules of the point form approach, the boost operator $U_{B(v)}$ is taken as a canonical one, obtaining that the transformed four-momenta are given by $p_{1,2}=B(v)k_{1,2}$ and satisfy the point form relation \begin{equation} \label{eq:pfe} p_1^\mu + p_2^\mu = \frac{P_N^\mu}{M_N} \left( \sqrt{\vec q\hspace{0.08cm}^2 + m_1^2} + \sqrt{\vec q\hspace{0.08cm}^2 + m_2^2} \right) \mbox{ }, \end{equation} where $P^\mu_N$ is the observed nucleon four-momentum and $M_N$ is its mass. It is worthwhile noting that Eq. (\ref{eqn:velocity-states}) redefines the single particle spins. Having applied canonical boosts, the conditions for a point form approach \cite{Klink:1998zz,melde} are satisfied. Therefore, the spins on the left hand state of Eq. (\ref{eqn:velocity-states}) perform the same Wigner rotations as $\vec k_1$ and $\vec k_2$, allowing to couple the spin and the orbital angular momentum as in the non relativistic case \cite{Klink:1998zz}, while the spins in the ket on the right hand of Eq. (\ref{eqn:velocity-states}) undergo the single particle Wigner rotations. In Point form dynamics, Eq. (\ref{eqn:H0}) corresponds to a good mass operator since it commutes with the Lorentz generators and with the four velocity. We diagonalize Eq. (\ref{eqn:H0}) in the Hilbert space spanned by the velocity states. Finally, instead of the internal momenta $\vec{k_1}$ and $\vec{k_2}$ we use the relative momentum $\vec q$, conjugate to the relative coordinate $\vec{r} = \vec{r}_1 - \vec{r}_2$, thus considering the following velocity basis states: \begin{equation} \vert v,\vec q,\lambda_1,\lambda_2 \rangle = U_{B(v)} \vert k_1,s_1,\lambda_1,k_2,s_2,\lambda_2 \rangle_{0} \mbox{ }. \end{equation} \begin{figure}[htbp] \centering \includegraphics[width=7cm]{fig1.eps} \caption{(Color online) Comparison between the calculated masses (black lines) of the $3^*$ and $4^*$ non strange baryon resonances (up to 2 GeV) and the experimental masses from PDG \cite{Nakamura:2010zzi} (boxes).} \label{fig:Spectrum3e4} \end{figure} \section{Results and discussion} Figure \ref{fig:Spectrum3e4} and Table \ref{tab:Spectrum} show the comparison between the experimental data \cite{Nakamura:2010zzi,Anisovich:2011fc} and the results of our quark-diquark model calculation, obtained with the set of parameters of Table \ref{tab:ResultingParameters}. In addition to the experimental data from PDG \cite{Nakamura:2010zzi}, we also consider the latest multi-channel Bonn-Gatchina partial wave analysis results, including data from Crystal Barrel/TAPS at ELSA and other laboratories \cite{Anisovich:2011fc}. In particular, these data differ from those of the PDG \cite{Nakamura:2010zzi} in the case of the $\Delta(1940) D_{33}$. \begin{table}[h1] \begin{center} \begin{tabular}{ccccccccc} \hline \hline \\ Resonance & Status & $M^{\mbox{exp.}}$ & $J^P$ & $L^P$ & $S$ & $s_1$ & $n_r$ & $M^{\mbox{calc.}}$ \\ & & (MeV) & & & & & & (MeV) \\ \\ \hline \\ $N(939)$ $P_{11}$ & **** & 939 & $\frac{1}{2}^+$ & $0^+$ & $\frac{1}{2}$ & 0,1 & 0 & 939 \\ $N(1440)$ $P_{11}$ & **** & 1420 - 1470 & $\frac{1}{2}^+$ & $0^+$ & $\frac{1}{2}$ & 0,1 & 1 & 1412 \\ $N(1520)$ $D_{13}$ & **** & 1515 - 1525 & $\frac{3}{2}^-$ & $1^-$ & $\frac{1}{2}$ & 0,1 & 0 & 1533 \\ $N(1535)$ $S_{11}$ & **** & 1525 - 1545 & $\frac{1}{2}^-$ & $1^-$ & $\frac{1}{2}$ & 0,1 & 0 & 1533 \\ $N(1650)$ $S_{11}$ & **** & 1645 - 1670 & $\frac{1}{2}^-$ & $1^-$ & $\frac{3}{2}$ & 1 & 0 & 1667 \\ $N(1675)$ $D_{15}$ & **** & 1670 - 1680 & $\frac{5}{2}^-$ & $1^-$ & $\frac{3}{2}$ & 1 & 0 & 1667 \\ $N(1680)$ $F_{15}$ & **** & 1680 - 1690 & $\frac{5}{2}^+$ & $2^+$ & $\frac{1}{2}$ & 0,1 & 0 & 1694 \\ $N(1700)$ $D_{13}$ & *** & 1650 - 1750 & $\frac{3}{2}^-$ & $1^-$ & $\frac{3}{2}$ & 1 & 0 & 1667 \\ $N(1710)$ $P_{11}$ & *** & 1680 - 1740 & $\frac{1}{2}^+$ & $0^+$ & $\frac{1}{2}$ & 0,1 & 2 & 1639 \\ $N(1720)$ $P_{13}$ & **** & 1700 - 1750 & $\frac{3}{2}^+$ & $2^+$ & $\frac{1}{2}$ & 0,1 & 0 & 1694 \\ $N(1875)$ $D_{13}$ & *** & 1820 - 1920 & $\frac{3}{2}^-$ & $1^-$ & $\frac{1}{2}$ & 0,1 & 1 & 1866 \\ $N(1880)$ $P_{11}$ & ** & 1835 - 1905 & $\frac{1}{2}^+$ & $0^+$ & $\frac{1}{2}$ & 0,1 & 3 & 1786 \\ $N(1895)$ $S_{11}$ & ** & 1880 - 1910 & $\frac{1}{2}^-$ & $1^-$ & $\frac{1}{2}$ & 0,1 & 1 & 1866 \\ $N(1900)$ $P_{13}$ & *** & 1875 - 1935 & $\frac{3}{2}^+$ & $0^+$ & $\frac{3}{2}$ & 0 & 0 & 1780 \\ missing & -- & -- & $\frac{3}{2}^+$ & $2^+$ & $\frac{1}{2}$ & 0,1 & 1 & 1990 \\ $N(2000)$ $F_{15}$ & ** & 1950 - 2150 & $\frac{5}{2}^+$ & $2^+$ & $\frac{1}{2}$ & 0,1 & 1 & 1990 \\ \\ $\Delta(1232)$ $P_{33}$ & **** & 1230 - 1234 & $\frac{3}{2}^+$ & $0^+$ & $\frac{3}{2}$ & 1 & 0 & 1236 \\ $\Delta(1600)$ $P_{33}$ & *** & 1500 - 1700 & $\frac{3}{2}^+$ & $0^+$ & $\frac{3}{2}$ & 1 & 1 & 1687 \\ $\Delta(1620)$ $S_{31}$ & **** & 1600 - 1660 & $\frac{1}{2}^-$ & $1^-$ & $\frac{1}{2}$ & 1 & 0 & 1600 \\ $\Delta(1700)$ $D_{33}$ & **** & 1670 - 1750 & $\frac{3}{2}^-$ & $1^-$ & $\frac{1}{2}$ & 1 & 0 & 1600 \\ $\Delta(1750)$ $P_{31}$ & * & 1708 - 1780 & $\frac{1}{2}^+$ & $0^+$ & $\frac{1}{2}$ & 1 & 0 & 1857 \\ $\Delta(1900)$ $S_{31}$ & ** & 1840 - 1920 & $\frac{1}{2}^-$ & $1^-$ & $\frac{1}{2}$ & 1 & 1 & 1963 \\ $\Delta(1905)$ $F_{35}$ & **** & 1855 - 1910 & $\frac{5}{2}^+$ & $2^+$ & $\frac{3}{2}$ & 1 & 0 & 1958 \\ $\Delta(1910)$ $P_{31}$ & **** & 1860 - 1920 & $\frac{1}{2}^+$ & $2^+$ & $\frac{3}{2}$ & 1 & 0 & 1958 \\ $\Delta(1920)$ $P_{33}$ & *** & 1900 - 1970 & $\frac{3}{2}^+$ & $2^+$ & $\frac{3}{2}$ & 1 & 0 & 1958 \\ $\Delta(1930)$ $D_{35}$ & *** & 1900 - 2000 & $\frac{5}{2}^-$ & $1^-$ & $\frac{3}{2}$ & 1 & 0 & 2064 \\ $\Delta(1940)$ $D_{33}$ & ** & 1940 - 2060 & $\frac{3}{2}^-$ & $1^-$ & $\frac{1}{2}$ & 1 & 1 & 1963 \\ $\Delta(1950)$ $F_{37}$ & **** & 1915 - 1950 & $\frac{7}{2}^+$ & $2^+$ & $\frac{3}{2}$ & 1 & 0 & 1958 \\ \\ \hline \hline \end{tabular} \end{center} \caption{Comparison between the experimental \cite{Nakamura:2010zzi} values of non strange baryon resonances masses (up to 2 GeV) and the numerical ones (all values are expressed in $MeV$). Tentative assignments of $2^*$ and $1^*$ resonances are shown in the second part of the table. $J^P$ and $L^P$ are respectively the total angular momentum and the orbital angular momentum of the baryon, including the parity $P$; $S$ is the total spin, obtained coupling the spin of the diquark $s_1$ and that of the quark; finally $n_r$ is the number of nodes in the radial wave function.} \label{tab:Spectrum} \end{table} \begin{table} \begin{center} \begin{tabular}{ccc} \hline \hline & & \\ $m_S$ $\mbox{(MeV)}~~$ & $m_{AV} - m_S$ $\mbox{(MeV)}~~$ & Source \\ & & \\ \hline \\ 730 & 210 & Bloch {\it et al.} \cite{Bloch:1999ke} \\ 750$\div$860 & 10$\div$170 & Oettel {\it et al.} \cite{Oettel:2002wf} \\ - & 290 & Wilczek \cite{Wilczek:2004im} \\ - & 210 & Jaffe \cite{Jaffe:2004ph} \\ 600 & 350 & Ferretti {\it et al.} \cite{Ferretti:2011zz} \\ 852 & 224 & Galata and Santopinto \cite{Galata:2012xt} \\ - & 200$\div$300 & Lichtenberg {\it et al.} \cite{Lichtenberg:1979de} \\ 770 & 140 & de Castro {\it et al.} \cite{deCastro:1993sr} \\ 420 & 520 & Sch\"{a}fer {\it et al.} \cite{Schafer:1993ra} \\ 692 & 330 & Cahill {\it et al.} \cite{Cahill:1995ka} \\ 595 & 205 & Lichtenberg {\it et al.} \cite{Lichtenberg:1996fi} \\ 737 & 212 & Burden {\it et al.} \cite{Burden:1996nh} \\ 688 & 202 & Maris \cite{Maris} \\ - & 360 & Orginos \cite{Orginos:2005vr} \\ 750 & 100 & Flambaum {\it et al.} \cite{Flambaum:2005kc} \\ 590 & 210 & \\ - & 162 & Babich {\it et al.} \cite{Babich:2007ah} \\ - & 270 & Eichmann {\it et al.} \cite{Eichmann:2008ef} \\ 740 & 210 & Hecht {\it et al.} \cite{Hecht:2002ej} \\ - & 135 & Santopinto and Galata \cite{Santopinto:2011mk} \\ 710 & 199 & Ebert {\it et al.} \cite{Ebert} \\ -- & 183 & Chakrabarti {\it et al.} \cite{Chakrabarti:2010zz} \\ 780 & 280 & Roberts {\it et al.} \cite{Roberts:2011cf} \\ 150 & 210 & This work \\ \\ \hline \hline \end{tabular} \end{center} \caption{Mass difference (in MeV) between scalar and axial-vector diquarks, according to some previous studies.} \label{tab:massediquarkaltri} \end{table} The spin-isospin transition interaction of Eq. (\ref{eqn:Vtr(r)}) mixes quark-scalar diquark and quark-axial-vector diquark states, i.e. states with $s_1 = 0$ ($t_1 = 0$) and $s_1 = 1$ ($t_1 = 1$), whose total spin (isospin) is $S = \frac{1}{2}$ ($T = \frac{1}{2}$). Thus, in this version of the model the nucleon state, as well as states such as the $D_{13}(1520)$, the $S_{11}(1535)$ and the $P_{11}(1440)$, contains both a $s_1 = 0$ and a $s_1 = 1$ component. In particular, the nucleon state, obtained by solving the eigenvalue problem of Eq. (\ref{eqn:H0}), in a schematic notation can be written as: \begin{equation} \label{eqn:nucleon.state} \left| N \right\rangle = 0.727 \left| qD_S, L=0 \right\rangle + 0.687 \left| qD_{AV}, L=0 \right\rangle \mbox{ }, \end{equation} where $D_S$ and $D_{AV}$ stand for the scalar and axial-vector diquarks, respectively, and $q$ for the quark. The radial wave functions (in momentum space) of the quark-scalar diquark [$\Phi_S(q)$] and quark-axial-vector diquark [$\Phi_{AV}(q)$] systems of Eq. (\ref{eqn:nucleon.state}), obtained by solving the eigenvalue problem of Eq. (\ref{eqn:H0}), can be fitted by harmonic oscillator wave functions \begin{subequations} \begin{equation} \Phi_S(q) = \frac{2 \alpha_S^{3/2}}{\pi^{1/4}} \mbox{ } e^{-\frac{1}{2} \alpha_S^2 q^2} \mbox{ }, \end{equation} \begin{equation} \Phi_{AV}(q) = \frac{2 \alpha_{AV}^{3/2}}{\pi^{1/4}} \mbox{ } e^{-\frac{1}{2} \alpha_{AV}^2 q^2} \mbox{ }, \end{equation} \end{subequations} with $\alpha_S = 3.29$ GeV$^{-1}$ and $\alpha_{AV} = 3.04$ GeV$^{-1}$. This parametrization can then be used to compute observables, such as the nucleon electromagnetic form factors. The introduction of the interaction of Eq. (\ref{eqn:Vtr(r)}) determines an improvement in the overall quality of the reproduction of the experimental data (considering only $3^*$ and $4^*$ resonances), with respect to that obtained with the previous version of this model \cite{Ferretti:2011zz}. In particular, the Roper resonance, $N(1440)$ $P_{11}$, is far better reproduced than before and the same holds for $N(1680)$ $F_{15}$. The present version of the relativistic quark-diquark model predicts only one missing state below the energy of 2 GeV (see Tab. \ref{tab:Spectrum}), while three quarks QM's give rise to several missing states \cite{Nakamura:2010zzi}. For example, Capstick and Isgur's model \cite{Capstick:1986bm} has 5 missing states up to 2 GeV, the hypercentral QM \cite{Giannini:2001kb} has 8, Glozman and Riska's model has 4 \cite{Glozman-Riska} and the U(7) model has 17 \cite{Bijker:1994yr}. The only missing resonance of our model, $N \frac{3}{2}^+(1990)$, lies at the same energy of the three star state $N(2000)$ $F_{15}$, which was previously a two star state of the PDG \cite{Nakamura:2010zzi}. Indeed the two resonances, $N \frac{3}{2}^+(1990)$ and $N(2000)$ $F_{15}$, have the same quantum numbers, except for the total angular momentum, because their spin ($\frac{1}{2}$) and orbital angular momentum (2) are coupled to $J^P = \frac{3}{2}^+$ or $\frac{5}{2}^+$. Thus, to split the two resonances one should take a spin-orbit interaction into account. While the absolute values of the diquark masses are model dependent, their difference is not. Comparing our result for the mass difference $m_{AV} - m_S$ between the axial-vector and the scalar diquark to those reported in Tab. \ref{tab:massediquarkaltri}, it is interesting to note that our estimation is comparable with all the others. Such evaluations come from phenomenological observations \cite{Jaffe:2004ph,Wilczek:2004im,Lichtenberg:1996fi}, lattice QCD calculations \cite{Orginos:2005vr,Babich:2007ah}, instanton liquid model calculations \cite{Schafer:1993ra}, applications of Dyson-Schwinger, Bethe-Salpeter and Fadde'ev equations \cite{Burden:1996nh,Hecht:2002ej,Maris,Cahill:1995ka,Flambaum:2005kc,Bloch:1999ke,Eichmann:2008ef} and constituent quark-diquark model calculations \cite{Ferretti:2011zz,Lichtenberg:1979de,deCastro:1993sr,Santopinto:2011mk}. The whole mass operator of Eq. (\ref{eqn:H0}) is diagonalized by means of a numerical variational procedure, based on harmonic oscillator trial wave functions. With a variational basis made of $N = 200$ harmonic oscillator shells, the results converge very well. We think that the present paper can be helpful to the experimentalists in their analysis of the properties of the $N$ and $\Delta$-type resonances. Our quark-diquark model results may be compared to those of three quarks QM's, showing a larger number of missing resonances. Our results may then help the experimentalists to distinguish between the two interpretations for baryons. Finally, in the future we will use our quark-diquark model wave functions to compute the nucleon electromagnetic form factors and the helicity amplitudes of baryon resonances \cite{DeSanctis:TBP}.
\section{Introduction} \label{intro} There is a great deal of interest in the distribution of spin and orbital angular momentum in hadronic systems. In general the underlying dynamics of partons in hadrons is relativistic. In the relativistic case the coefficients that relate the parton spins to hadronic spins are momentum-dependent and are necessarily influenced by the momentum-dependence of the hadronic wave function. This momentum dependence appears in both relativistic quantum mechanics and relativistic quantum field theory. In addition, in QCD, because constituent partons are confined, their ``mass'' becomes an additional dynamical variable. The purpose to this paper is to show how this dynamical dependence enters the relation between the hadronic and partonic spins and to show the equivalence of the treatment spin in Poincar\'e and Lorentz covariant theories. The treatment of spin in relativistic systems is different than it is in non-relativistic systems. In a relativistic system the spin of a parton is identified with the angular momentum of the parton in its rest frame while the spin of the hadron is defined as the angular momentum of the hadron in its rest frame. Transforming a parton from its rest frame to the hadrons rest frame, where the spins can be coupled, involves boosts which generate dynamical rotations. These rotations transform the parton spins and also impact the relative orbital angular momentum before they can be coupled. The spin of the constituents, the internal orbital angular momentum and the spin of the system are related by Clebsch-Gordan coefficients of the Poincar\'e group\cite{moussa}\cite{bkwp}\cite{wgwp} . The Poincar\'e group Clebsch-Gordan coefficients are labeled by eigenvalues of mass and spin Casimir operators, which are dynamical operators. An additional complication is that boosts to the rest frame are not unique; a boost to the rest frame followed by a momentum-dependent rotation is a different boost to the rest frame. There are as many different kinds of boosts as there are momentum-dependent rotations. Each boost defines a different spin observable. For example, there are distinct boosts that are used to define the helicity, light front spin or canonical spin. These are three specific choices, that are distinguished by useful properties, out of an infinite number of possibilities. In many-body systems there is another relevant spin, which we call the constituent spin\cite{wgwp}, which is distinguished by the property that spins and orbital angular momenta can be combined using ordinary SU(2) coupling methods to get the hadronic spin. While all of the spins satisfy $SU(2)$ commutation relations, the different spins observables are related by the momentum dependent (Melosh) rotations\cite{melosh} that relate different boosts. Because different spin observables differ by momentum-dependent rotations, partial derivatives with respect to momentum that hold one spin observable constant will not commute with a different spin observable. This means not only are there an infinite number of possible spin observables, but each one is associated with a different quantity that can be identified with an orbital angular momentum. As a result the spin and orbital angular momentum content of a hadron is dynamical and representation dependent. In what follows we discuss some of the relevant issues. The dependence of the spin on the choice of boost is seen in the relations between the spin and angular momentum \begin{equation} j_a^l := {1 \over 2} \epsilon^{lmn} {B_a^{-1} (p)^m{}_{\mu}} {B_a^{-1} (p)^n{}_{\nu}}J^{\mu \nu} \label{eq:a1} \end{equation} where $B_a^{-1} (p)^m{}_{\mu}$ is a boost that maps $p$ to its rest frame. Here $a$ is an index used to distinguish different types of Lorentz boosts. This definition can be equivalently expressed in terms of the polarization vectors, $e_a^{m}{}_{\mu}(p):= B_a^{-1} (p)^m{}_{\mu}$, $m=1,2,3$, that are three orthonormal space-like vectors that are orthogonal to the four momentum: \begin{equation} j_a^l = {1 \over 2} \epsilon^{lmn} {e_a^{m}{}_{\mu}(p) e_a^{n}{}_{\nu}(p)} J^{\mu \nu} . \label{eq:a2} \end{equation} Spins constructed using different boosts (labeled by $a$ and $b$) are related by momentum-dependent Melosh rotations \begin{equation} j_a^l = R_{ab}(p)^l{}_m j_b^m \qquad \mbox{where} \qquad R_{ab}(p)^l{}_m= B_a^{-1} (p)^l{}_{\mu} B_b (p)^{\mu}_{m} . \label{eq:a3} \end{equation} Because the different types of spin observables differ by momentum-dependent rotations, ``Position operators'' \cite{newton}\cite{wp}\cite{bkwp}that involve partial derivatives with respect to momentum need to specify which kind of spins is being held constant during the differentiation, \begin{equation} [\pmb{\nabla}_{P_{\vert_{j_a}}}, \mathbf{j}_a]=0 \, {\Rightarrow} \, [\pmb{\nabla}_{P_{\vert_{j_a}}}, \mathbf{j}_b] \not= 0 . \label{eq:a4} \end{equation} These partial derivatives can be written in terms of the Poincar\'e generators with $\mathbf{V}= \mathbf{P}/M$ by \cite{wp}\cite{bkwp} \begin{equation} X^k_a= {i \pmb{\nabla}_{\mathbf{P}_{\vert_{\mathbf{j}_a}}}} = -{1 \over 2} \{H^{-1},{K}^k\} + i H^{-1} C_{1a}^{kl}(\mathbf{V}) j_a^l \label{eq:xx} \end{equation} where $H$ is the Hamiltonian, $M= \sqrt{H^2-\mathbf{P}^2}$ is the invariant mass operator and $\mathbf{V}$ is the four velocity. In terms of these operators the spins and angular momentum are related by \begin{equation} J^j = (\mathbf{X}_a \times \mathbf{P})^j + C_{2a}^{jk} (\mathbf{V})j_a^k \label{eq:a5} \end{equation} where the operators $C_{1a}^{jk}(\mathbf{V})$ and $ C_{2a}^{jk} (\mathbf{\mathbf{V}})$ are the following functions of the Poincar\'e generators: \begin{equation} C_{1a}^{jk}(\mathbf{V}) = {1 \over 2} \mathbf{Tr} [B_a(\mathbf{V})^{-1}\sigma_j B_a(\mathbf{V}) \sigma_k] - V^0 \mathbf{Tr} [B_a(\mathbf{V})^{-1}{\partial \over \partial V_l} B_a(\mathbf{V}) \sigma_m] \label{eq:a6} \end{equation} \begin{equation} C_{2a}^{jk} ({\mathbf{V}}) = {1 \over 2} \mathbf{Tr} [B_a(\mathbf{V})^{-1}\sigma_j B_a(\mathbf{V}) \sigma_k] + i \epsilon_{jlm} \mathbf{Tr} [B_a(\mathbf{V})^{-1}{\partial \over \partial V_l} B_a(\mathbf{V}) \sigma_m] \label{eq:a7} \end{equation} and $B_a(\mathbf{V})$ is the $SL(2,\mathbb{C})$ representation of the $a$-boost. The quantity $\mathbf{X}_a \times \mathbf{P}$ is the associated orbital angular momentum \cite{wp}\cite{bkwp} Three components of the four momentum and the projection of any of these spin observables on a given axis are labels for vectors in irreducible subspaces. Products of two such irreducible representations can be expressed as direct integrals of composite irreducible representations using the Clebsch-Gordan coefficients for the Poincar\'e group. Like any set of Clebsch-Gordan coefficients, the actual coefficients depend on the choice of irreducible basis. The Poincar\'e group Clebsch-Gordan coefficients for a basis labeled by the $a$-type spin are \[ _a\langle (M_1,j_1) \mathbf{P}_1 ,\mu_1 (M_2,j_2) \mathbf{P}_2 ,\mu_2 \vert k, j (M_1,j_1,M_2,j_2) \mathbf{P} , \mu , l , s_{12} \rangle_a = \] \[ \sum_{\mu_1',\mu_2',\mu_1'',\mu_2'',\mu_s,m} \delta (\mathbf{P} - \mathbf{P}_1-\mathbf{P}_2) {\delta (k -k(\mathbf{P}_1,\mathbf{P}_2)) \over k^2} \times \] \[ \sqrt{{\omega_{M_1} (\mathbf{k})\omega_{M_2} (\mathbf{k}) \over \omega_{M_1} (\mathbf{P}_1)\omega_{M_2} (\mathbf{P}_2)}} \sqrt{{ \omega_{M_1} (\mathbf{P}_1)+\omega_{M_2} (\mathbf{P}_1) \over \omega_{M_1} (\mathbf{k})+\omega_{M_2} (\mathbf{k})}} \times \] \[ {D^{j_1}_{\mu_1 \mu_1'}[R_{wa} (B_a(V) ,k_1)] D^{j_1}_{\mu_1'\mu_1''} [R_{ac}(k_1)]} \times \] \[ {D^{j_2}_{\mu_2 \mu_2'}[R_{wa} (B_a(V) ,k_2)] D^{j_2}_{\mu_2'\mu_2''} [R_{ac}(k_2)]} \times \] \begin{equation} Y^{l}_{m} (\hat{\mathbf{k}}(\mathbf{P}_1,\mathbf{P}_2)) \langle j_1, \mu_1'', j_2, \mu_2'' \vert s_{12}, \mu_s \rangle \langle l, m, s_{12}, \mu_s \vert j, \mu \rangle \label{eq:a8} \end{equation} These involve two types of spin rotations. There are Wigner rotations $R_{wa} (B_a(V) ,k_i)$ that arise from the $a$-boosts that relate the system and parton rest frames and generalized Melosh rotations, $R_{ac}(k_2)$, that transform the resulting spins to the canonical spin representation where all of the spins and orbital angular momenta Wigner rotate together so they can be added using ordinary $SU(2)$ spin addition. The spins obtained by applying these two rotations to the hadronic spins are the constituent spins mentioned earlier. These are the spins associated with the magnetic quantum numbers $\mu_i''$ in (\ref{eq:a8}). It is apparent from this equation that when these spins are combined with the orbital angular momentum using spherical harmonics and SU(2) Clebsch-Gordan coefficients the result is the total spin. The Poincar\'e group Clebsch-Gordan coefficients (\ref{eq:a8}) simplify in special bases. If the spins are defined using the standard rotationless boosts there are no Melosh rotations, if the rotationless boost is replaced by a light-front boost there are no Wigner rotations, and if the rotationless boost is replaced by a helicity boost the Wigner rotations become multiplication by a phase. One result of the momentum-dependence of the rotations is that the momentum-dependence of the hadronic wave function affects the expectation values of both the spins and orbital angular momentum. Since the momentum dependence of the hadronic wave function is determined by the dynamics, the dynamics enters in the spin coupling when the Poincar\'e Clebsch-Gordan coefficients are integrated against the hadronic wave functions. When one couples two interacting subsystems, one has to ask whether the masses in the Poincar\'e Clebsch-Gordan coefficients are the physical masses of the subsystems or the invariant masses of their constituents. For example, the mass of a meson or the invariant mass of a quark antiquark pair? So far we have treated them as invariant masses of the constituents. Cluster properties suggest that one should really use the physical mass operators of the subsystems. Fortunately there is a unitary transformation that removes the interaction dependence from the hadronic spin\cite{fcwp}\cite{wp2}. In this representation the spins can be coupled by sequential coupling using the Clebsch-Gordan coefficients of the Poincar\'e group as if the particles were not interacting. This unitary transformation changes the Hamiltonian, generating many-body interactions. It also changes the representation of the wave function in a way that preserves probabilities, expectation values, as well as scattering observables. As a result of this all of the dynamical spin effects can be absorbed by changing the representation of the wave function. In QCD the quark masses themselves are also not constant. A second ambiguity with spin has to do with whether the dynamics is formulated using Poincar\'e covariant or Lorentz covariant bases. Field theories are normally formulated using Lorentz covaraint bases while relativistic quantum mechanics is typically formulated using Poincar\'e covariant bases. These are simply related; the dynamics enters both representations, but in different but equivalent ways. To understand this note that the unitary representation of the Poincar\'e group on positive-mass positive-energy irreducible basis states has the form \begin{equation} U(\Lambda ,0) \vert (M,j)\mathbf{P},\mu \rangle_a = \sum \vert (M,j)\pmb{\Lambda}P,\nu \rangle_a D^j_{\nu \mu}(R_{wa}(\Lambda,P)). \label{eq:a9} \end{equation} The Wigner rotation can be decomposed into the composition of a boost followed by a Lorentz transformation followed by an inverse boost with the transformed four momentum \begin{equation} R_{wa}(\Lambda,P) = B_a^{-1}(\Lambda P) \Lambda B_a (P). \label{eq:a10} \end{equation} The group representation property can be used to split the Wigner function apart. The finite dimensional representations of $SU(2)$ are related to finite dimensional representation of $SL(2,\mathbb{C})$ by analytic continuation\cite{weinberg}\cite{wgwp}, so we can still use the group representation property. Absorbing the Wigner functions of the boosts into the states gives the Lorentz spinor representation of the states: \begin{equation} \vert (m,j) \mathbf{P}, {b} \rangle := \sum_{\mu} \vert (m,j) \mathbf{P}, \mu \rangle_a D^j_{\mu {b}}[B_a^{-1}(P/M)]. \label{eq:a11} \end{equation} Here the boosts are represented by $2 \times 2 \, SL(2,\mathbb{C})$ transformations. These spinor basis states (\ref{eq:a11}) have the following Lorentz covariant transformation property \begin{equation} {U(\Lambda, 0)} \vert (m,j) \mathbf{P}, {b} \rangle = \sum_{{b}'} \vert (m,j) \pmb{\Lambda}{P}, {b}' \rangle {D^j_{{b}'{b}}[\Lambda]}. \label{eq:a12} \end{equation} The price paid for using the covariant representation is that the Hilbert space inner product becomes dynamical \[ \langle \psi \vert \phi \rangle = \] \begin{equation} \int \langle \psi \vert (m,j) \mathbf{P}, {b} \rangle d^4P \theta(P^0) \delta (P^2 + {M^2}) D^j_{bb'}[P^{\mu}\sigma_{\mu}/{M}] \langle (m,j) \mathbf{P}, {b}' \vert \phi \rangle \label{eq:a13} \end{equation} where we have used the hermiticity of the $SL(2,\mathbb{C})$ representation of the rotationless boost which gives \begin{equation} B_a (V)B_a^{\dagger}(V) = B_c (V)R_{ca}(V) R^{\dagger}_{ca} B_c^{\dagger}(V) = B_c (V)B_c^{\dagger}(V) = B^2_c (V) = P^{\mu}\sigma_{\mu}/M \label{eq:a14} \end{equation} independent of the type $(a)$ of boost. The zero component of $\sigma^{\mu}$ in (\ref{eq:a14}) is the identity and the other three components are the Pauli matrices. In (\ref{eq:a13}) the dynamics is appears in the mass-shell condition, which makes the Wigner function into a positive matrix. The inner product (\ref{eq:a13}) is identical to the original Poincar\'e covariant inner product. Unlike representations of $SU(2)$, the representations of $SL(2,C)$ are not equivalent to the complex conjugate representations. This means that we could alternatively replace (\ref{eq:a11}) by \begin{equation} \vert (m,j) \mathbf{P}, \dot{b} \rangle := \sum_{\mu} \vert (m,j) \mathbf{P}, \mu \rangle_a D^j_{\mu \dot{b}}[B_a^{\dagger}(P/M)]. \label{eq:a11b} \end{equation} and (\ref{eq:a12}) by \begin{equation} {U(\Lambda, 0)} \vert (m,j) \mathbf{P}, \dot{b} \rangle = \sum_{{b}'} \vert (m,j) \pmb{\Lambda}{P}, \dot{b}' \rangle {D^j_{\dot{b}'\dot{b}}[((\Lambda)^{\dagger})^-1)]}. \label{eq:a12b} \end{equation} This gives a representation of the scalar product that has the same form as (\ref{eq:a13}) with the replacement \begin{equation} D^j_{bb'}[P^{\mu}\sigma_{\mu}/{M}] \to D^{j}_{\dot{b} \dot{b}'}[P^{\mu} \sigma_2\sigma^*_{\mu}\sigma_2/{M} ] . \label{eq:a16} \end{equation} While the inner products in all three representations are identical, the Lorentz covariant and its complex conjugate representations are related by space reflection. Space reflection changes the kernel of the Hilbert-space scalar product in the covariant representations. Space reflection can be represented as an operator on states by replacing the representations (\ref{eq:a11}) and (\ref{eq:a11b}) by a direct sum of both representations. In the direct sum representation the wave function becomes a $2 \times (2j+1)$ component spinor \begin{equation} \psi ({P} , b) \to \left ( \begin{array}{c} \xi ({P} , b) \\ \chi ({P} , \dot{b}) \end{array} \right ) \label{eq:a17} \end{equation} and the kernel of the inner product becomes \begin{equation} d^4P \theta(P^0) \delta (P^2 + {M^2}) \left ( \begin{array}{cc} D^j_{bb'}[P^{\mu}\sigma_{\mu}/{M}] & 0 \\ 0 & D^{j}_{\dot{b} \dot{b}'}[P^{\mu} \sigma_2\sigma^*_{\mu}\sigma_2/{M} ] \end{array} \right ). \label{eq:a18} \end{equation}. One desirable feature of the Lorentz covariant representation is that the basis-dependent features are hidden in the wave functions. To see this note that for rotations the upper and lower components have identical transformations laws \begin{equation} U(R, 0) \vert (M,j) \mathbf{P}, {b} \rangle = \sum_{\dot{b}'} \vert (M,j) R\mathbf{P}, {b}' \rangle D^j_{{b}'{b}}[R]. \label{eq:a19} \end{equation} and \begin{equation} U(R, 0) \vert (M,j) \mathbf{P}, \dot{b} \rangle = \sum_{\dot{b}'} \vert (M,j) R\mathbf{P}, \dot{b}' \rangle D^j_{\dot{b}'\dot{b}}[R]. \label{eq:a20} \end{equation} which is the standard rotational transformation law that leads to the standard relation \begin{equation} \mathbf{J} = \mathbf{X} \times \mathbf{P} + \mathbf{j} \label{eq:a21} \end{equation} in the covariant representation. In this representation the relation between the spin, angular momentum, and orbital angular momentum looks very much like the corresponding non-relativistic quantities. The price paid for this simplification is that the Hilbert space inner product has a non-trivial kernel. This kernel contains all of the dynamical effects discussed in the context of Poincar\'e irreducible spins. The covariant spin is related to the Poincar\'e irreducible spin of a particle by a boost. For spin $1/2$ particles the usual $u$ and $v$ spinors are direct sum representations of a Lorentz boost. The choice (helicity, canonical, light front spin) appears in the representation of these spinors. The Poincar\'e irreducible labels are the parameters that normally label asymptotic states in the $S$-matrix. In the end there are many different kinds of spin observables. In order to measure the spin we need to know how the various spin operators couple to the electroweak current operators. This will be different for each type of spin observable. The conclusion is that for relativistic systems the coupling of spins and orbital angular momenta involves momentum and mass-dependent Poincar\'e group Clebsch-Gordan coefficients. In QCD even the parton masses that appear in these coefficients become variables. The result is that the dynamics cannot be ignored in attempts to identify the different terms the contribute to the hadronic spin. In addition, there are many different spin and orbital angular momentum observables. How these different quantities contribute to the hadronic spin is representation dependent. As a consequence, is important to know how these operators are precisely defined and how they are related to experimental observables. This research was supported by the US DOE Office of Science, under grant number No. DE-FG02-86ER40286
\section*{Table of contents} \begin{itemize} \footnotesize \item[{\scriptsize (p.~\pageref*{pdf:Bilen})}] \hyperlink{Bilen.1}{C. Bilen (INRIA Rennes, France), Srdan Kitic (INRIA Rennes, France), N. Bertin (IRISA CNRS, France), R. Gribonval (INRIA Rennes, France),\newline ``Sparse Acoustic Source Localization with Blind Calibration for Unknown Medium Characteristics''.} \item[{\scriptsize (p.~\pageref*{pdf:Boumal})}] \hyperlink{Boumal.1}{N. Boumal (UCL, Belgium), B. Mishra (ULg, Belgium), P.-A. Absil (UCL, Belgium), R. Sepulchre (Cambridge U., UK),\newline ``Manopt: a Matlab toolbox for optimization on manifolds''.} \item[{\scriptsize (p.~\pageref*{pdf:Bundervoet})}] \hyperlink{Bundervoet.1}{S. Bundervoet (VUB-ETRO,Belgium), C. Schretter (VUB-ETRO,Belgium), A.Dooms (VUB-ETRO,Belgium), P. Schelkens (VUB-ETRO, Belgium),\newline ``Bayesian Estimation of Sparse Smooth Speckle Shape Models for Motion Tracking in Medical Ultrasound''.} \item[{\scriptsize (p.~\pageref*{pdf:Chabiron})}] \hyperlink{Chabiron.1}{O. Chabiron (U.Toulouse,France), F.Malgouyres (U.Toulouse,France), J.-Y.,Tourneret (U.Toulouse,France), N. Dobigeon (U. Toulouse, France),\newline ``Learning a fast transform with a dictionary''.} \item[{\scriptsize (p.~\pageref*{pdf:Chainais})}] \hyperlink{Chainais.1}{P. Chainais (LAGIS / INRIA Lille, France), C. Richard (Lab. Lagrange, Nice, France),\newline ``A diffusion strategy for distributed dictionary learning''.} \item[{\scriptsize (p.~\pageref*{pdf:Cornelis})}] \hyperlink{Cornelis.1}{B. Cornelis (VUB-ETRO, Belgium), A. Dooms (VUB-ETRO, Belgium), I. Daubechies (Duke U., USA), D. Dunson (Duke U., USA),\newline ``Bayesian crack detection in high resolution data''.} \item[{\scriptsize (p.~\pageref*{pdf:Dankova})}] \hyperlink{Dankova.1}{M. Dankova (Brno University of Technology) and P. Rajmic (Brno University of Technology, Czech Republic) \newline ``Compressed sensing of perfusion MRI''.} \item[{\scriptsize (p.~\pageref*{pdf:Degraux})}] \hyperlink{Degraux.1}{K. Degraux (UCL, Belgium), V. Cambareri (U.Bologna, Italy), B.Geelen (IMEC, Belgium), L. Jacques (UCL, Belgium), G. Lafruit (IMEC, Belgium), G. Setti (U. Bologna, Italy),\newline ``Compressive Hyperspectral Imaging by Out-of-Focus Modulations and Fabry-P\'erot Spectral Filters''.} \item[{\scriptsize (p.~\pageref*{pdf:Determe})}] \hyperlink{Determe.1}{J.-F. Determe (ULB, Belgium), J. Louveaux (UCL, Belgium), F. Horlin (ULB, Belgium),\newline ``Filtered Orthogonal Matching Pursuit: Applications''.} \item[{\scriptsize (p.~\pageref*{pdf:Dremeau})}] \hyperlink{Dremeau.1}{A. Dr\'emeau (ESPCI ParisTech, France), P. H\'eas (INRIA, Rennes, France), C. Herzet (INRIA, Rennes, France),\newline ``Combining sparsity and dynamics: an efficient way''.} \item[{\scriptsize (p.~\pageref*{pdf:Duval})}] \hyperlink{Duval.1}{V. Duval (U. Paris-Dauphine), G. Peyr\'e (U. Paris-Dauphine),\newline ``Discrete vs. Continuous Sparse Regularization''.} \item[{\scriptsize (p.~\pageref*{pdf:Fawzi})}] \hyperlink{Fawzi.1}{A. Fawzi (LTS4, EPFL, Switzerland), M. Davies (U. Edinburgh, UK), P. Frossard (LTS4, EPFL, Switzerland),\newline ``Dictionary learning for efficient classification based on soft-thresholding''.} \item[{\scriptsize (p.~\pageref*{pdf:Gillis})}] \hyperlink{Gillis.1}{N. Gillis (U. Mons, Belgium), S. A. Vavasis (U. Waterloo, UK),\newline ``Semidefinite Programming Based Preconditioning for More Robust Near-Separable Nonnegative Matrix Factorization''.} \item[{\scriptsize (p.~\pageref*{pdf:Herzet})}] \hyperlink{Herzet.1}{C. Herzet (INRIA Rennes, France), C. Soussen (U. de Lorraine, Nancy, France),\newline ``Enhanced Recovery Conditions for OMP/OLS by Exploiting both Coherence and Decay''.} \item[{\scriptsize (p.~\pageref*{pdf:Kitic})}] \hyperlink{Kitic.1}{S. Kitic, (Inria, France), N. Bertin, (IRISA, France), R. Gribonval (Inria, France),\newline ``Wideband Audio Declipping by Cosparse Hard Thresholding''.} \item[{\scriptsize (p.~\pageref*{pdf:LeMagoarou})}] \hyperlink{LeMagoarou.1}{L. Le Magoarou (INRIA, Rennes, France), R. Gribonval (INRIA, Rennes, France),\newline ``Strategies to learn computationally efficient and compact dictionaries''.} \item[{\scriptsize (p.~\pageref*{pdf:Liang})}] \hyperlink{Liang.1}{J. Liang (GREYC, ENSICAEN, France), J. Fadili (GREYC, ENSICAEN, France), G. Peyr\'e (U. Paris-Dauphine, France),\newline ``Iteration-Complexity for Inexact Proximal Splitting Algorithms''.} \item[{\scriptsize (p.~\pageref*{pdf:Liutkus})}] \hyperlink{Liutkus.1}{A. Liutkus (Institut Langevin, France), D. Martina (Institut Langevin, France), S. Gigan (Institut Langevin, France), L. Daudet (Institut Langevin, France),\newline ``Calibration and imaging through scattering media''.} \item[{\scriptsize (p.~\pageref*{pdf:Maggioni})}] \hyperlink{Maggioni.1}{M. Maggioni (Duke U.), S. Minsker (Duke U.), N. Strawn (Duke U.),\newline ``Multiscale Dictionary and Manifold Learning: Non-Asymptotic Bounds for the Geometric Multi-Resolution Analysis''.} \item[{\scriptsize (p.~\pageref*{pdf:Mory})}] \hyperlink{Mory.1}{C. Mory (iMagX Belgium and CREATIS U. Lyon France) and L. Jacques (UCLouvain, Belgium) \newline ``An application of the Chambolle-Pock algorithm to 3D + time tomography''.} \item[{\scriptsize (p.~\pageref*{pdf:Ngole})}] \hyperlink{Ngole.1}{F. Ngole (CEA-CNRS-Paris 7, France), Starck, Jean-Luc (CEA-CNRS-Paris 7, France),\newline ``Super-resolution method using sparse regularization for point spread function recovery''.} \item[{\scriptsize (p.~\pageref*{pdf:Schretter})}] \hyperlink{Schretter.1}{C. Schretter (VUB-ETRO, Belgium), I. Loris (ULB, Belgium), A. Doom (VUB-ETRO, Belgium), P. Schelkens (VUB- ETRO, Belgium),\newline ``Total Variation Reconstruction From Quasi-Random Samples''.} \item[{\scriptsize (p.~\pageref*{pdf:Vaiter})}] \hyperlink{Vaiter.1}{S. Vaiter (U. Paris-Dauphine, France), M. Golbabaee (U. Paris-Dauphine, France), J. Fadili (GREYC, ENSICAEN, France), G. Peyr\'e (U. Paris-Dauphine, France),\newline ``Model Selection with Piecewise Regular Gauges''.} \item[{\scriptsize (p.~\pageref*{pdf:Vukobratovic})}] \hyperlink{Vukobratovic.1}{D. Vukobratovic (U. of Novi Sad, Serbia), A. Pizurica (Ghent University, Belgium),\newline ``Adaptive Compressed Sensing Using Sparse Measurement Matrices''.} \end{itemize} \newpage \setcounter{page}{1} \renewcommand{\thepage}{\arabic{page}} \label{pdf:Bilen} \includepdf[offset=1cm -4mm,pages=-,link=true,linkname=Bilen,pagecommand={}]{Bilen.pdf} \label{pdf:Boumal} \includepdf[offset=1cm -4mm,pages=-,link=true,linkname=Boumal,pagecommand={}]{Boumal.pdf} \label{pdf:Bundervoet} \includepdf[offset=1cm -4mm,pages=-,link=true,linkname=Bundervoet,pagecommand={}]{Bundervoet.pdf} \label{pdf:Chabiron} \includepdf[offset=1cm -4mm,pages=-,link=true,linkname=Chabiron,pagecommand={}]{Chabiron.pdf} \label{pdf:Chainais} \includepdf[offset=1cm -4mm,pages=-,link=true,linkname=Chainais,pagecommand={}]{Chainais.pdf} \label{pdf:Cornelis} \includepdf[offset=1cm -4mm,pages=-,link=true,linkname=Cornelis,pagecommand={}]{Cornelis.pdf} \label{pdf:Dankova} \includepdf[offset=1cm -4mm,pages=-,link=true,linkname=Dankova,pagecommand={}]{Dankova.pdf} \label{pdf:Degraux} \includepdf[offset=1cm -4mm,pages=-,link=true,linkname=Degraux,pagecommand={}]{Degraux.pdf} \label{pdf:Determe} \includepdf[offset=1cm -4mm,pages=-,link=true,linkname=Determe,pagecommand={}]{Determe.pdf} \label{pdf:Dremeau} \includepdf[offset=1cm -4mm,pages=-,link=true,linkname=Dremeau,pagecommand={}]{Dremeau.pdf} \label{pdf:Duval} \includepdf[offset=1cm -4mm,pages=-,link=true,linkname=Duval,pagecommand={}]{Duval.pdf} \label{pdf:Fawzi} \includepdf[offset=1cm -4mm,pages=-,link=true,linkname=Fawzi,pagecommand={}]{Fawzi.pdf} \label{pdf:Gillis} \includepdf[offset=1cm -4mm,pages=-,link=true,linkname=Gillis,pagecommand={}]{Gillis.pdf} \label{pdf:Herzet} \includepdf[offset=1cm -4mm,pages=-,link=true,linkname=Herzet,pagecommand={}]{Herzet.pdf} \label{pdf:Kitic} \includepdf[offset=1cm -4mm,pages=-,link=true,linkname=Kitic,pagecommand={}]{Kitic.pdf} \label{pdf:LeMagoarou} \includepdf[offset=1cm -4mm,pages=-,link=true,linkname=LeMagoarou,pagecommand={}]{LeMagoarou.pdf} \label{pdf:Liang} \includepdf[offset=1cm -4mm,pages=-,link=true,linkname=Liang,pagecommand={}]{Liang.pdf} \label{pdf:Liutkus} \includepdf[offset=1cm -4mm,pages=-,link=true,linkname=Liutkus,pagecommand={}]{Liutkus.pdf} \label{pdf:Maggioni} \includepdf[offset=1cm -4mm,pages=-,link=true,linkname=Maggioni,pagecommand={}]{Maggioni.pdf} \label{pdf:Mory} \includepdf[offset=1cm -4mm,pages=-,link=true,linkname=Mory,pagecommand={}]{Mory.pdf} \label{pdf:Ngole} \includepdf[offset=1cm -4mm,pages=-,link=true,linkname=Ngole,pagecommand={}]{Ngole.pdf} \label{pdf:Schretter} \includepdf[offset=1cm -4mm,pages=-,link=true,linkname=Schretter,pagecommand={}]{Schretter.pdf} \label{pdf:Vaiter} \includepdf[offset=1cm -4mm,pages=-,link=true,linkname=Vaiter,pagecommand={}]{Vaiter.pdf} \label{pdf:Vukobratovic} \includepdf[offset=1cm -4mm,pages=-,link=true,linkname=Vukobratovic,pagecommand={}]{Vukobratovic.pdf} \newpage \null \vfill \hfill Finished on \today, Louvain-la-Neuve, Belgium. \null \end{document}
\section{Introduction} We consider secure message transmission over a wiretap channel $W:{\mathcal X}\rightarrow {\mathcal Y} \times {\mathcal Z}$ with noiseless, public feedback. For each transmission $x\in {\mathcal X}$ over $W$, the receiver observes a random output $Y \in {\mathcal Y}$ and an eavesdropper observes a correlated side-information $Z\in {\mathcal Z}$, with probability $W(Y, Z|x)$. Furthermore, the receiver can send a feedback to the transmitter over a noiseless channel. However, the feedback channel is public and any communication sent over it is available to the eavesdropper. The transmitter seeks to send a message $M$ to the receiver without revealing it to the eavesdropper. For a given probability of error $\epsilon$ and a given secrecy parameter $\delta$, what is the maximum possible rate $C_{\ep, \delta}$ of a transmitted message? For a degraded wiretap channel $W$ with no feedback, the wiretap capacity $C = \inf_{\epsilon, \delta}C_{\ep, \delta}$ was established in the seminal work of Wyner \cite{Wyn75ii} where it was shown that \[ C = \max_{\bPP X}\CMI XYZ. \] The capacity of a general wiretap channel was established in \cite{CsiKor78}. Extensions to wiretap channels with general statistics were considered in \cite{Hay06}. The model with feedback considered here was introduced in \cite{Cheong76} where it was noted that the availability of a noiseless feedback can enable positive rates of transmission over a wiretap channel with zero capacity (see, also, \cite{Mau93}). However, the wiretap capacity with feedback remains unknown in general; $\max_{\bPP X}\CMI XYZ$ constitutes an upper bound on it. In this paper, we establish a {\it strong version} of this bound and show that for $\epsilon+\delta<1$ \[ C_{\ep, \delta} \le \max_{\bPP X}\CMI XYZ, \] thereby characterizing $C_{\ep, \delta}$ for all $0<\epsilon,\delta<1$ for a degraded wiretap channel. A partial strong converse for a degraded wiretap channel was established in \cite{MorganW14} for a restricted range of $\epsilon, \delta$. Another strong converse for a degraded wiretap channel for the case when $\delta\rightarrow 0$ was established, concurrently to this work, in \cite{TanB14}. In this work, we show a strong converse for all values of $\epsilon$ and $\delta$. Our proof relies on a slight modification of a recent reduction of hypothesis testing to secret key agreement shown in \cite{TyaWat14, TyaWat14ii}. Specifically, we show that a wiretap channel code yields an active hypothesis test for distinguishing between two channels \cite{Hay09ii}. Consequently, the rate of a wiretap code is bounded above by the rate of the optimum exponent of the probability of error of type II for discriminating a channel $W$ from another channel $V$ such that $V(y,z|x) = V_2(z|x)V_1(y|z)$, given that the probability of error of type I is less than $\epsilon+\delta$. This gives an upper bound on the length of a wiretap code, which leads to the strong converse upon using the characterization of the optimal exponent for channel discrimination derived in \cite{Hay09ii}. This approach is along the lines of {\it meta-converse} of \cite{PolPooVer10}, where a reduction of hypothesis testing to channel coding was used to establish a finite-blocklength converse for the channel coding problem (see, also, \cite{Nagaoka01} and \cite[Section 4.6]{Hayashi06}). Our main result is given in the next section. Section \ref{s:hypothesis_testing} and \ref{s:secret_key} contains a review of relevant results in binary hypothesis testing and secret key agreement, respectively. The final section contains a proof of our main result. \section{Main result}\label{s:main_result} We describe a generalization of the classic wiretap channel coding problem \cite{Wyn75ii, CsiKor78} that was considered in \cite{Cheong76, Mau93, AhlCsi93}, where, in addition to transmitting over the wiretap channel, the terminals can communicate using a noiseless, public feedback channel from the receiver to the transmitter. A wiretap code for a discrete\footnote{The restriction to discrete alphabet is cosmetic. Our results apply to channels with continuous alphabet. In particular, our strong converse holds for the Gaussian wiretap channel \cite{CheHel78}.} memoryless wiretap channel $W:{\mathcal X}\rightarrow {\mathcal Y}\times {\mathcal Z}$ with feedback consists of (possibly randomized) encoder mappings $e_t: \{1, ...,N\} \times {\mathcal F}^{t} \rightarrow {\mathcal X}$, $1\le t\le n$, feedback mappings $f_t: {\mathcal Y}^t \rightarrow {\mathcal F}$, $0\le t\le n-1$, and a decoder $d: {\mathcal Y}^n \rightarrow \{1, ..., N\}$. For a random message $M \sim \mathtt{unif}\{1,..., N\}$, the protocol begins with a feedback $F_0$ from the receiver at $t=0$. Subsequently, at each time instance $1\le t\le n-1$ the transmitter sends $X_t = e_t(M, F^{t-1})$ and the channel outputs $(Y_t, Z_t)$ with probability $W(Y_t, Z_t| X_t)$. The receiver observes $Y_t$ and sends feedback $F_t = f_t(Y^t)$, and the eavesdropper observes $Z_t$. The protocol stops with a final transmission $X_n = e_n(M, F^{n-1})$ over the channel and the subsequent decoding $\hat M = d(Y^n)$ by the receiver. We denote by $\mathbf{F}$ the overall feedback communication $F_0, ..., F_{n-1}$. The mappings $(\{e_t\}_{t=1}^n, \{f_t\}_{t=0}^{n-1}, d)$ constitute an $(N,n, \epsilon, \delta)$ wiretap code if \[ \bPr {M\neq \hat M}\le \epsilon, \] and \[ \ttlvrn {\bPP {MZ^n \mathbf{F}}}{\bPP M\times \bPP{Z^n\mathbf{F}}}\le \delta, \] where $\ttlvrn \mathrm{P} \mathrm{Q}$ denotes the variation distance between $\mathrm{P}$ and $\mathrm{Q}$ given by \[ \ttlvrn \mathrm{P} \mathrm{Q} = \frac 12 \sum_x|\mathrm{P}(x) - \mathrm{Q}(x)|. \] A rate $R>0$ is $(\ep, \delta)$-achievable if there exists an $(\lfloor 2^{nR}\rfloor, n ,\epsilon, \delta)$ wiretap code for all $n$ sufficiently large. The $(\ep, \delta)$-{\it wiretap capacity} $C_{\ep, \delta}$ is the supremum of all $(\ep, \delta)$-achievable rates. Our main result in an upper bound on $C_{\ep, \delta}$ \begin{theorem}\label{t:main_result} For $0\le \epsilon, \delta$ with $\epsilon+\delta< 1$, the $(\ep, \delta)$-wiretap capacity is bounded above as \[ C_{\ep, \delta} \le \max_{\bPP X}\CMI XYZ. \] \end{theorem} For the special case of a degraded wiretap channel $W$ with $W(y,z|x) = W_1(y|x)W_2(z|y)$, Theorem \ref{t:main_result} yields a {\it strong converse} for wiretap capacity. \begin{corollary} For a degraded wiretap channel $W$, \[ C_{\ep, \delta} = \begin{cases} \displaystyle \max_{\bPP X} \CMI XYZ, &\quad 0< \epsilon < 1-\delta, \\ \displaystyle \max_{\bPP X} \MI XY, &\quad 1-\delta \le \epsilon <1. \end{cases} \] \end{corollary} {\it Proof.} For $0< \epsilon <1-\delta$, the result is an immediate corollary of Theorem \ref{t:main_result} and \cite{Wyn75ii}\footnote{While the secrecy criterion in \cite{Wyn75ii} is different from variational secrecy required here, the achievability result for the latter follows from the results in \cite{Csi96, Hay06}.}. For $1-\delta\le \epsilon <1$, the converse follows from the strong converse for the capacity of a DMC with feedback ({\it cf.}{} \cite{PolVer10}). Moving to the proof of achievability, it suffices to restrict to $\epsilon+\delta=1$. For this case, achievability follows by randomizing between an $(\epsilon_n, 1)$ wiretap code, $\epsilon_n \rightarrow 0$ as $n\rightarrow \infty$, and a $(1, 0)$ wiretap code -- the randomizing bit is communicated as the public feedback $F_0$ by the receiver\footnote{Alternatively, the sender can transmit the randomizing bit over the wiretap channel with neglible rate loss.} \qed As a preparation for the proof of Theorem \ref{t:main_result} given in Section~\ref{s:proof_main_result}, we review some results in hypothesis testing and secret key agreement in the next two sections. \section{Hypothesis testing}\label{s:hypothesis_testing} Consider a simple binary hypothesis testing problem with null hypothesis $\mathrm{P}$ and alternative hypothesis $\mathrm{Q}$, where $\mathrm{P}$ and $\mathrm{Q}$ are distributions on the same alphabet ${\cal X}$. Upon observing a value $x\in {\mathcal X}$, the observer needs to decide if the value was generated by the distribution $\bPP{}$ or the distribution $\mathrm{Q}$. To this end, the observer applies a stochastic test $\mathrm{T}$, which is a conditional distribution on $\{0,1\}$ given an observation $x\in {\mathcal X}$. When $x\in {\mathcal X}$ is observed, the test $\mathrm{T}$ chooses the null hypothesis with probability $\mathrm{T}(0|x)$ and the alternative hypothesis with probability $T(1|x) = 1 - T(0|x)$. For $0\leq \epsilon<1$, denote by $\beta_\epsilon(\mathrm{P},\mathrm{Q})$ the infimum of the probability of error of type II given that the probability of error of type I is less than $\epsilon$, i.e., \begin{eqnarray} \beta_\epsilon(\mathrm{P},\mathrm{Q}) := \inf_{\mathrm{T}\, :\, \mathrm{P}[\mathrm{T}] \ge 1 - \epsilon} \mathrm{Q}[\mathrm{T}], \nonumber \end{eqnarray} where \begin{eqnarray*} \mathrm{P}[\mathrm{T}] &=& \sum_x \mathrm{P}(x) \mathrm{T}(0|x), \\ \mathrm{Q}[\mathrm{T}] &=& \sum_x \mathrm{Q}(x) \mathrm{T}(0|x). \end{eqnarray*} The following result credited to Stein characterizes the optimum exponent of $\beta_\epsilon(\mathrm{P}^n,\mathrm{Q}^n)$ where $\mathrm{P}^n = \mathrm{P} \times ...\times \mathrm{P}$ and $\mathrm{Q}^n = \mathrm{Q}\times ... \times \mathrm{Q}$. \begin{lemma}\emph{({\it cf.}{} \cite[Theorem 3.3]{Kul68})} For every $0< \epsilon <1$, we have \begin{eqnarray} \lim_{n\to\infty} - \frac{1}{n} \log \beta_\epsilon(\mathrm{P}^n,\mathrm{Q}^n) = D(\mathrm{P} \| \mathrm{Q}), \nonumber \end{eqnarray} where $D(\mathrm{P}\|\mathrm{Q})$ is the Kullback-Leibler divergence given by \begin{eqnarray*} D(\mathrm{P}\|\mathrm{Q}) = \sum_{x\in {\mathcal X}} \mathrm{P}(x) \log \frac{\mathrm{P}(x)}{\mathrm{Q}(x)}, \end{eqnarray*} with the convention $0\log(0/0) = 0$. \end{lemma} Next, we review a problem of active hypothesis testing where the distribution at each instance is determined by a prior action. Specifically, given two DMCs $W:{\mathcal X}\rightarrow {\mathcal Y}$ and $ V:{\mathcal X}\rightarrow {\mathcal Y}$, we seek to design a transmission-feedback scheme such that by observing the channel inputs, channel outputs, and feedback we can determine if the underlying channel is $W$ or $V$. Formally, an $n$-length active hypothesis test consist of (possibly randomized) encoder mappings $e_t:{\mathcal F}^t\rightarrow {\mathcal X}$, $1\le t\le n$, feedback mappings $f_t:{\mathcal Y}^t \rightarrow {\mathcal F}$, $0\le t\le n-1$, and a conditional distribution $T$ on $\{0,1\}$ given $X^n, Y^n, \mathbf{F}$. On observing $X^n, Y^n, \mathbf{F}$, we detect the null hypothesis $W$ with probability $T(0| X^n, Y^n, \mathbf{F})$ and alternative hypothesis $V$ with probability $T(1| X^n, Y^n, \mathbf{F})$. Analogous to $\beta_\epsilon(\mathrm{P}, \mathrm{Q})$, the quantity $\beta_\epsilon(W,V,n)$, for $0\le \epsilon<1$, is the infimum of the probability of error of type II over all $n$ length active hypothesis tests for null hypothesis $W$ and alternative hypothesis $V$ such that the probability of error of type I is no more than $\epsilon$. The following analogue of Stein's lemma for active hypothesis testing was established in \cite{Hay09ii} (see, also, \cite{PolVer10}). \begin{theorem}[\cite{Hay09ii}]\label{t:active_HT} For $0< \epsilon <1$, \begin{align*} \lim_n - \frac 1n \log \beta_\ep(W,V,n) &= \max_{\bPP X} \CKL WV{\bPP X} \\ &= \max_x \KL {W_x}{ V_x}, \end{align*} where $W_x$ and $V_x$, respectively, denote the $x$th row of $W$ and $V$. \end{theorem} Remarkably, the exponent above is achieved without any feedback, {\it i.e.}, while feedback is available, it does not help to improve the asymptotic exponent of $\beta_\ep(W,V,n)$. \section{Secret key agreement}\label{s:secret_key} In this section, we review two party secret key (SK) agreement where parties observing random variables $X$ and $Y$ communicate interactively over a public channel to agree on a SK that is concealed from an eavesdropper with access to the communication and a side-information $Z$. Formally, the parties communicate using an interactive communication $\mathbf{F} = F_1, ..., F_r$ where $F_1 = F_1(X), F_2 = F_2(Y, F_1), F_3 = F_3(X, F^2)$, $F_4 = F_4(Y, F^3)$ and so on. A random variable $K = K(X,\mathbf{F})$ constitutes an $(\ep, \delta)$-SK if there exists $\hat K = \hat K(Y, \mathbf{F})$ such that \begin{align} \bPr {K\neq \hat K}\le \epsilon, \nonumber \end{align} and \begin{align} \ttlvrn {\bPP {KZ\mathbf{F}}}{\bPP {\mathtt{unif}} \times \bPP {Z\mathbf{F}}}\le \delta. \nonumber \end{align} The following upper bound on the number of values $k$ taken by an $(\ep, \delta)$-SK $K$ was shown in \cite{TyaWat14, TyaWat14ii}: \[ \log k \le - \log \beta_{\ep+\delta+\eta}(\bPP{XYZ}, \bQQ{XYZ}) + 2\log \frac 1\eta, \] for all $0< \eta < 1-\epsilon -\delta$, and all $\bQQ{XYZ} = \bQQ{X|Z} \bQQ{Y|Z} \bQQ{Z}$. Underlying the proof of this bound is an intermediate reduction argument in \cite[Lemma 1]{TyaWat14} that relates SK agreement to hypothesis testing. We recall this result below. \begin{theorem}[\cite{TyaWat14, TyaWat14ii} ]\label{t:SK_reduction} For $0\le \epsilon, \delta, \epsilon+\delta<1$, let random variables $K,\hat K$, and $Z$ be such that $\bPr {K \neq \hat K}\le \epsilon$ and \[ \ttlvrn {\bPP{KZ}}{\bPP {\mathtt{unif}}\times\bPP{Z }}\le \delta, \] where $\bPP {\mathtt{unif}}$ denotes a uniform distribution on $k$ values. Then, for every $0<\eta< 1-\epsilon -\delta$ and every $\bQQ {K\hat K Z} = \bQQ{K\mid Z}\bQQ{\hat K \mid Z}\bQQ Z$, \[ \log k \le - \log \beta_{\ep+\delta+\eta}(\bPP{K\hat K Z},\bQQ{K \hat KZ}) + 2\log \frac 1\eta. \] \end{theorem} \section{Proof of main result}\label{s:proof_main_result} We present a converse result that applies for every fixed $n$ and is asymptotically tight, giving the strong converse result of Theorem \ref{t:main_result}. \begin{theorem}\label{t:wiretap_converse} For $0\le \epsilon, \delta, \epsilon+\delta< 1$, given an $(N, n, \epsilon, \delta)$-wiretap code, we have \[ \log N \le -\log \beta_{\ep+\delta+\eta}(W,V,n) + 2\log \frac 1\eta, \] for all $0< \eta < 1-\epsilon-\delta$ and all channels $V: {\mathcal X} \rightarrow {\mathcal Y} \times {\mathcal Z}$ such that $V(y, z|x) = V_2(z|x)V_1(y|z)$. \end{theorem} {\it Proof of Theorem \ref{t:main_result}.} Theorem \ref{t:main_result} follows form Theorems \ref{t:wiretap_converse} and \ref{t:active_HT} upon noting that for $W(y, z|x) = W_2(z|x)W_1(y|z, x)$ \begin{align*} &\hspace{-1em}\min_V \max_{\bPP X}\,\CKL W V {\bPP X} \\ &= \min_{V_1}\max_{\bPP X}\, \CKL {W_1}{V_1}{\bPP X W_2} \\ &= \max_{\bPP X} \min_{V_1}\,\CKL {W_1}{V_1}{\bPP X W_2} \\ &= \max_{\bPP X}\, \CKL {\bPP {Y\mid ZX}}{\bPP{Y\mid Z}}{\bPP{ZX}} \\ &= \max_{\bPP X}\,\CMI XYZ, \end{align*} where $\bPP{XYZ}$ is given by $\bPP XW$. \qed We need the following result to prove Theorem~\ref{t:wiretap_converse}. \begin{lemma}\label{l:conditional_independence} For a wiretap channel $V:{\mathcal X}\rightarrow {\mathcal Y} \times {\mathcal Z}$ such that $V(y,z|x) = V_2(z|x)V_1(y|z)$, a random message $M$, and a wiretap code, let $\hat M =d(Y^n)$ and $\mathbf{F}$ be the corresponding feedback. Then, the induced distribution $\bQQ{M \hat M Z^n \mathbf{F}}$ satisfies factorization condition \[ \bQQ{M \hat M\mid Z^n \mathbf{F}} = \bQQ{M\mid Z^n\mathbf{F}}\times\bQQ{\hat M\mid Z^n \mathbf{F}}. \] \end{lemma} {\it Proof of Lemma~\ref{l:conditional_independence}.} Denote by $U_x$ and $U_y$, respectively, the local randomness at the transmitter and the receiver, and by $F^t$ the feedback $(F_0, ..., F^t)$. Thus, the encoder mapping $e_t$ is a (deterministic) function of $(M, U_x, F^{t-1})$ and the feedback mapping $f_t$ is a (deterministic) function of $(Y^t, U_y)$. The proof entails a repeated application of the fact that conditionally independent random variables remain so when conditioned additionally on an interactive communication (cf.~\cite{TyaNar13ii}) and is completed by induction. Specifically, note first that $\bQQ{MU_xU_y \mid F_0} = \bQQ{MU_x\mid F_0}\bQQ{U_y \mid F_0}$ since $(M,U_x)$ and $U_y$ are independent and $F_0$ is an interactive communication. Under the induction hypothesis \begin{align*} &\hspace{-1em}\bQQ{MU_xX^{t-1}U_yY^{t-1}\mid Z^{t-1}F^{t-1}} \\ &=\bQQ{MU_xX^{t-1}\mid Z^{t-1} F^{t-1}} \bQQ{U_yY^{t-1}\mid Z^{t-1} F^{t-1}}, \end{align*} we get \begin{align*} &\hspace{-1em}\CMI{M,U_x,X^t}{U_y,Y^t}{Z^t, F^{t-1}} \\ &= \CMI{M,U_x,X^t}{U_y,Y^{t-1}}{Z^t, F^{t-1}} \\ &\leq \CMI{M,U_x,X^t}{U_y,Y^{t-1}}{Z^{t-1}, F^{t-1}} \\ &= \CMI{M,U_x,X^{t-1}}{U_y,Y^{t-1}}{Z^{t-1}, F^{t-1}} \\ &=0, \end{align*} where the first equality and inequality follow since $Y_t$ and $Z_t$, respectively, are outputs of $V_1$ for input $Z_t$ and $V_2$ for input $X_t$, and the second equality holds since $X_t = e_t(M, U_x, F^{t-1})$, which completes the proof. \qed {\it Proof of Theorem \ref{t:wiretap_converse}.} Given an $(N, n ,\epsilon, \delta)$ wiretap code, a message $M\sim \mathtt{unif}\{1,...,N\}$ and its decoded value $\hat M = d(Y^n)$ satisfy the conditions for Theorem~\ref{t:SK_reduction} with $K = M, \hat K = \hat M,$ and $Z = (Z^n, \mathbf{F})$. Letting $\bQQ {M\hat M Z^n\mathbf{F}}$ be the distribution on $(M,\hat M, Z^n, \mathbf{F})$ when the underlying channel is $V$, by Lemma~\ref{l:conditional_independence} and Theorem~\ref{t:SK_reduction} we get \[ \log N \le - \log \beta_{\ep+\delta+\eta}(\bPP {M\hat M Z^n\mathbf{F}}, \bQQ {M\hat M Z^n\mathbf{F}}) + 2\log \frac 1\eta. \] Note that a test for the simple binary hypothesis testing problem for $\bPP {M\hat M Z^n\mathbf{F}}$ and $\bQQ {M\hat M Z^n\mathbf{F}}$ along with the wiretap code constitutes an active hypothesis test for $W$ and $V$. Therefore, \begin{align*} &\hspace*{-1em}- \log \beta_{\ep+\delta+\eta}(\bPP {M\hat M Z^n\mathbf{F}}, \bQQ {M\hat M Z^n\mathbf{F}}) \\ &\le - \log \beta_{\ep+\delta+\eta}(W,V,n), \end{align*} which completes the proof.\qed \section*{Acknowledgements} MH is partially supported by a MEXT Grant-in-Aid for Scientific Research (A) No. 23246071. MH is also partially supported by the National Institute of Information and Communication Technology (NICT), Japan. The Centre for Quantum Technologies is funded by the Singapore Ministry of Education and the National Research Foundation as part of the Research Centres of Excellence programme.
\section{Introduction} Since Worthington's pioneering work for the splashing of fluid impact~\cite{Worthington}, the impact splashing has been studied intensively by many physicists. Recent developments of the numerical computation and the high-speed imaging enable us to reveal the detailed dynamics of the fluid impact~\cite{Yarin,Clanet,ThoroddsenRev}. Particularly, a liquid droplet impact either onto a liquid pool or a hard floor has been studied well so far~\cite{Yarin}. A solid sphere impact to a non-Newtonian target fluid has been also investigated experimentally~\cite{Tabuteau,Ara}. Furthermore, some recent studies have concerned granular target impacts as well as fluid target impacts. Examples include a solid sphere's impact to a granular bed~\cite{Uehara,Walsh,Katsuragi2007,Katsuragi2013,Marston,Royer}, a liquid droplet impact to a granular bed~\cite{Katsuragi,Katsuragi2}, and a vortex ring impact to a granular bed~\cite{Sano}. In spite of these recent efforts for the soft matter impact, fingering instability by the impact between a solid sphere and a liquid pool has not been studied well. This is a little surprising because it is probably one of the simplest setups. So far, the study of fingering instability has been almost restricted within a droplet impact. However, natural fingering instability is not limited in the macroscopic droplet impact. For instance, microscopic impact cratering has been found in the space environment. Nakamura et al. have found wavy rim structures of tiny ($\simeq 100$ nm order) craters on the sample returned from the asteroid Itokawa~\cite{Nakamura}. This crater rim structure reminds us the fingering instability. The petal-like crater's rim structure has been also found in the droplet impact to a granular bed~\cite{Katsuragi,Katsuragi2}. From the high-speed video data taken in the experiments, we have confirmed that the petal-like structure is caused by the fingering instability of the impacting droplet. Petal-like ejecta blanket structure called Rampart crater can be also found in the Martian surface and a laboratory experiments of the hypervelocity impact to a muddy target~\cite{Rampart}. Such simple laboratory experiments might be useful to discuss the origin of widely variated natural craters shapes. Note that, however, surfaces of actual astronomical objects do not consist of liquid. They are rather covered with regolith. Moreover, Itokawa's returned samples, on which nano craters were found, are solid (not liquid). Thus the real craters morphologies might not directly link to the liquid impact. However, we believe that the fundamental study of the liquid impact could be a first step to understand the variety of complex craters morphologies. To classify the physical origins of various fingering-like structures in soft matter impacts, fundamental investigations on the fingering instability with various projectiles and targets are necessary. For example, viscoelastic property of the target material might play a certain role in the very high-speed impact in which the melting of the target could occur. As a first step, we should begin with the simplest setup. Therefore, we carry out one of the simplest impact experiments---a steel sphere impact onto a viscous liquid---in order to approach a fundamental law of the fingering instability. Particularly, we analyze the experimental data by combining the models of liquid deformation and Rayleigh-Taylor instability. Using the derived model, we speculate the impact velocity to make astronomical nano craters accompanied with wavy rim structure. Then the limitations for this consideration are also discussed finally. \section{Experimental} We build a simple experimental apparatus as schematically shown in figure~\ref{fig:f1}. A steel sphere in diameter $D_p=3$, $6.35$, or $8$ (mm) is held by an electromagnet. Then it is dropped from a certain height $h$ to commence a free fall. The free-fall height $h$ ranges from $20$ to $605$ (mm) corresponding to the impact velocity of $U=0.2-3.6$ (m/s). This free-fall drop system is basically identical to that used in our previous experiment on the agar gel impact~\cite{Ara}. Density of the steel sphere is $\rho_p=8.0 \times 10^3$ (kg/m$^3$). We employ silicone oil (Shin-Etsu Chemical Co., Ltd.) or distilled water as target fluids. The range of kinematic viscosity $\nu$ and surface tension $\gamma$ of target fluids are $0.65 \leq \nu \leq 200$ ($\times 10^{-6}$ m$^2$/s$(=$cSt)) and $16 \leq \gamma \leq 72$ ($\times 10^{-3}$ N/m), respectively. The density of fluid ranges $0.8 \leq \rho_f \leq 1.0$ ($\times 10^3$ kg/m$^3$). Namely, main control parameters are the impact inertia and the target viscosity. Target surface tension and density are varied slightly. Actually, we have used the viscoelastic fluid target as well. However, we could not observe any fingering instability for viscoelastic targets, at least in the range of current experimental condition. Much larger impact inertia is probably required to induce the fingering instability in the viscoelastic impact. Thus we are going to focus only on the viscous target case, in this study. Impact and splashing are captured by a high-speed camera (Photron SA-5) at $5,000$ fps. All the experiments are performed under the atmospheric pressure environment ($\simeq 101$ kPa). \begin{figure} \begin{center} \scalebox{1.15}[1.15]{\includegraphics{apparatus.eps}} \end{center} \caption{A schematic drawing of the experimental apparatus. A stainless steel sphere is dropped onto a viscous liquid pool by free fall. The splashing is taken by a high-speed camera.} \label{fig:f1} \end{figure} \section{Results and discussion} Typical snapshots of the experimental results are presented in figure~\ref{fig:f2}. When the target viscosity is large, any fingering instability cannot be observed. Such an example ($\nu=120$ cSt) is shown in figure~\ref{fig:f2}(a). To induce the fingering instability on a very viscous target, much larger impact inertia is necessary. On the other hand, we can confirm the clear fingering instability for an impact to a less viscous target (figure~\ref{fig:f2}(b); $\nu = 0.89$ cSt). When the impact inertia increases more, random splashing is induced as shown in figure~\ref{fig:f2}(c). In this regime, it is hard to count the number of fingers from the raw video data. Perhaps, this limit comes from the limitations of temporal and spatial resolutions of the images taken by the high-speed camera. Much higher resolution images might enable us to study this random regime. Since we are interested in the typical fingering structure, we restricted ourselves to the transient regime in which the clear fingering structure can be observed. \begin{figure} \begin{center} \scalebox{0.9}[0.9]{\includegraphics{raw.eps}} \end{center} \caption{Typical snapshots of the impacts: (a) no-fingering case ($d=8$ mm, $\nu= 120$ cSt, and $h=240$ mm); (b) fingering case ($d=6.35$ mm, $\nu=0.89$ cSt, and $h=120$ mm); (c) random splashing ($d=6.35$ mm, $\nu =0.65$ cSt, and $h=605$ mm).} \label{fig:f2} \end{figure} Number of fingers $N$ is counted from the snapshots (like figure~\ref{fig:f2}(b)) by eye. Since the fingers move, split, and merge, it is not so easy to count $N$. Moreover, the fingering structure is finally relaxed relatively in a short time. To see the initial instability, we count all of possible fingers at the early stage of the impact. We would like to characterize $N$ by the scaling analysis using some relevant dimensionless numbers. \begin{figure*} \begin{center} \scalebox{1}[1]{\includegraphics{Nf_Fr.eps}} \end{center} \caption{Scaling results of the counted fingers number $N$ by (a) impact Reynolds number $Re_I$, (b) droplet-deformation-included Rayleigh-Taylor instability, and (c) crater-deformation-based Rayleigh-Taylor instability. The solid and dashed lines in (a) indicate the scaling $N \sim {Re_I}^{1/3}$ and $N\sim {Re_I}^{3/4}$, respectively. The solid line in (b) represents $N= We^{1/2}Re^{1/4}/4\sqrt{3}$. The solid line in (c) corresponds to $N=(\rho_r Fr)^{1/4}We^{1/2}/3^{3/4}$. While these scaling laws capture the global trend of data behavior, the data dispersion is significantly large.} \label{fig:f3} \end{figure*} \subsection{Impact Reynolds number} First, the simple scaling analysis is tried by combining two possibly relevant dimensionless numbers: Weber number $We = \rho_f U^2 D_p/\gamma$ and Reynolds number $Re=UD_p/\nu$. Here we introduce the impact Reynolds number~\cite{Marmanis}. Impact Reynolds number has been derived by considering the representative length scale $L=[\nu (m_p/\gamma)^{1/2}]^{1/2}$, where $m_p$ is the sphere's mass. The characteristic timescale $t_{\gamma}=(m_p/\gamma)^{1/2}$ comes from the spring-like effect of the liquid surface tension. And $L$ indicates the length scale of momentum diffusion by $t_{\gamma}$ and $\nu$. Then, the impact Reynolds number $Re_I$ is defined as \begin{equation} Re_I = \frac{UL}{\nu} = We^{1/4}Re^{1/2}. \label{eq:Impact_Reynolds_numer} \end{equation} Marmanis and Thoroddsen have shown that the number of fingers by a droplet impacting to a hard floor can be scaled by $N\sim {Re_I}^{3/4}$~\cite{Marmanis}. In figure~\ref{fig:f3}(a), the counted result of $N$ as a function of $Re_I$ for the current experiment is displayed. Although the data considerably scatter due to the difficulty of fingers-counting, here we apply the data fitting to $Re_I$ scaling. Then, as shown in figure~\ref{fig:f3}(a), the data can be fitted by a scaling, \begin{equation} N \sim {Re_I}^{\alpha}. \label{eq:number_scaling} \end{equation} The obtained scaling exponent $\alpha=1/3$ is different from the previously obtained droplet impact experiment, $\alpha=3/4$~\cite{Marmanis}. Solid and dashed lines in Fgi.~\ref{fig:f3}(a) correspond to $N\sim{Re_I}^{1/3}$ and $N\sim{Re_I}^{3/4}$, respectively. This difference might come from the geometry difference. While $\alpha=3/4$ was obtained for the impacting droplet case, $\alpha=1/3$ corresponds to the case that a hard projectile impact to a liquid target. Moreover, the obtained scaling means very weak $We$ and $Re$ dependences: $N \sim We^{1/12}Re^{1/6}$. Such small exponents are not so useful for the scaling analysis. It is hard to discuss the underlying physical mechanism on the basis of these small exponents. The data can be fitted only because the model of Eq.~(\ref{eq:number_scaling}) has two free fitting parameters. \subsection{Rayleigh-Taylor instability} Next, Rayleigh-Taylor instability is considered as an alternative candidate to explain the fingering. Rayleigh-Taylor instability is induced at an interface of denser and lighter fluids when the denser one is accelerated toward the lighter one. The capillary-based characteristic wave number $k$ for the Rayleigh-Taylor instability can be computed as~\cite{ChandrasekharFluid,Piriz2006,Bret2011}, \begin{equation} k=\sqrt{\frac{(\rho_h - \rho_l)a}{3\gamma}}, \label{eq:RT_wavenumber} \end{equation} where $\rho_h$, $\rho_l$, and $a$ are the density of denser fluid, that of lighter fluid, and the applied acceleration, respectively. By assuming $\rho_f = \rho_h \gg \rho_l$ and $a \simeq U^2/D_p$, $k$ can be rewritten as $k=We^{1/2}/\sqrt{3}D_p$. If this instability is induced at the perimeter of a circle in diameter $D_c$, the number of fingers $N$ is written as, \begin{equation} N = \frac{D_c}{2\sqrt{3}D_p}We^{1/2}. \label{eq:RT_N} \end{equation} By considering a liquid droplet impact onto a hard floor, the maximally deformed droplet diameter $D_c$ at viscosity-dominant regime can be approximated by~\cite{Mundo1995,Pasandideh-Fard1996,Bhola}, \begin{equation} D_c \simeq \frac{1}{2} D_p Re^{1/4}. \label{eq:Bhola_diameter} \end{equation} Here, $D_p$ corresponds to the initial droplet diameter. Substituting Eq.~(\ref{eq:Bhola_diameter}) to Eq.~(\ref{eq:RT_N}), one can obtain a form of $N$ as~\cite{Bhola}, \begin{equation} N= \frac{1}{4\sqrt{3}}We^{1/2}Re^{1/4}. \label{eq:Bhola_N} \end{equation} It should be noticed that this scaling variable $We^{1/2}Re^{1/4}$ is different from $Re_I(= We^{1/4}Re^{1/2})$. The $We^{1/2}Re^{1/4}$ is sometimes called the splashing parameter. The scaling form (Eq.~(\ref{eq:Bhola_N})) is compared with the current experimental result in figure~\ref{fig:f3}(b). As can be seen, this scaling seems to correspond to the lower limit of the experimental data. Although the scaling in figure~\ref{fig:f3}(b) might look worse than that in figure~\ref{fig:f3}(a), note that the scaling of Eq.~(\ref{eq:Bhola_N}) does not include any fitting parameter. To improve the scaling more, here we consider the crater cavity model~\cite{Engel1966,Engel1967}. Let us assume that the projectile's kinetic energy $E_k$ is mainly consumed by the potential energy of the crater's cavity $E_{\rm cav}$. The cavity potential energy $E_{\rm cav}$ is calculated as $E_{\rm cav} = \pi \rho_f g D_c^4/64$~\cite{Engel1966}. By equating $E_{\rm cav}$ and the impacting kinetic energy $E_k=\pi \rho_p D_p^3 U^2/12$, the cavity diamater $D_{c}$ is obtained as, \begin{equation} D_{c} = 2\left( \frac{\rho_p D_p^3 U^2}{3\rho_f g} \right)^{1/4}=2D_p\left( \frac{\rho_p}{3\rho_f}Fr\right)^{1/4}. \label{eq:cavity_radius} \end{equation} Substituting Eq.~(\ref{eq:cavity_radius}) to Eq.~(\ref{eq:RT_N}), a scaling form is obtained as, \begin{equation} N= \left( 3^{-3} \rho_r Fr \right)^{1/4}We^{1/2}, \label{eq:Engel_N} \end{equation} where $Fr = U^2/gD_p$ is Froude number and $\rho_r={\rho_p}/{\rho_f}$ is the density ratio between a projectile and a target. This form looks similar to Eq.~(\ref{eq:Bhola_N}); $Re$ is replaced by $\rho_r Fr$ in terms of the scaling relation. While the dominant effect in Eq.~(\ref{eq:Bhola_N}) is viscosity, the gravity plays an essential role in Eq.~(\ref{eq:Engel_N}). This implies that the scaling of Eq.~(\ref{eq:Engel_N}) is valid for the gravity-dominant cratering regime. In figure~\ref{fig:f3}(c), the experimental data are compared with this scaling. The abscissa of figure~\ref{fig:f3}(c) comes from the right-hand side of Eq.~(\ref{eq:Engel_N}). Whereas the data still widely scatter, the scaling captures the global trend without any fitting parameter. Perhaps, the small projectile case ($D_p=3$ mm) can be better explained by the viscosity-based scaling (Eq.~(\ref{eq:Bhola_N})). Other data seem to agree with the gravity-dominant scaling. This fingering relates to the surface tension effect and we did not directly measure the surface tension value (i.e. we use catalogue data). Therefore the data contain large uncertainty. That is probably the reason why it is hard to reduce the relatively large data dispersion in all panels of figure~\ref{fig:f3}. Moreover, the range of $We$ swept by this study is not wide enough. To make the scaling relation sure, more systematic experiments with widely ranging dimensionless numbers must be carried out. \subsection{Application to the fingering of nano craters found on the asteroid Itokawa} In principle, we can estimate the impact conditions for various (gravity-dominant) craters with fingering instability, by using Eq.~(\ref{eq:Engel_N}). However, it is not easy to estimate the proper value of $Fr$ particularly for the astronomical impacts. Instead, the crater dimensions are directly accessible as far as the craters are observed. In such a situation, Eq.~(\ref{eq:RT_N}) is useful to evaluate the impact conditions. This scaling is based only on the Rayleigh-Taylor instability. It does not assume the dominant dynamics of the deformation: e.g. viscosity, gravity, or strength. Here, let us estimate the impact velocity $U$ for the wavy rim structure of nano craters found on Itokawa's returned samples~\cite{Nakamura}. To this purpose, Eq.~(\ref{eq:RT_N}) is rewritten as, \begin{equation} U=\frac{2N}{D_c} \sqrt{\frac{3D_p \gamma}{\rho_f}}. \label{eq:U_RT} \end{equation} In order to compute a specific value of $U$, here we use physical properties of SiO$_2$: $\rho_f =2000$ kg/m$^3$ and $\gamma=0.014$ N/m~\cite{Blum}. In addition, $N\simeq 7$ and $D_c\simeq 200$ nm can be observed from the photos of nano craters~\cite{Nakamura}. Assuming $D_p=D_c/2$, which is a reasonable assumption for fluid impacts, we finally obtain $U \simeq 100$ m/s. If we assume a molten glass state for the target, $\gamma$ value is about one order of magnitude greater than the currently assumed value. However, the resultant $U$ still remains in the order of $U = 10^2$ m/s. The assumption $D_p=D_c/2$ actually means $Fr\simeq 1$ if we use Eq.~(\ref{eq:cavity_radius}). However, it is quite difficult to satisfy this condition under the very weak gravity like Itokawa. Thus, the crater diameter $D_c$ must be determined rather by the target strength $Y$. By considering the energy balance between $E_{\rm cav,Y}=\pi D_c^3 T/12$ and $E_k$, $D_c$ is written as, \begin{equation} D_c =D_p\left(\frac{\rho_p U^2}{Y}\right). \label{eq:strength_Dc} \end{equation} Then, the strength $Y=2.5$ MPa is obtained from the condition $D_p=D_c/2$. We can even estimate the finite strength of the target in this framework. The existence of the finite strength implies that the target material should somehow be viscoplastic. Substituting Eq.~(\ref{eq:strength_Dc}) to Eq.~(\ref{eq:U_RT}), we obtain a form of the impact velocity in strength-dominant cratering regime as, \begin{equation} U =\left( \frac{12 N^2 \gamma}{\rho_f D_p}\right)^{3/10}\left(\frac{Y}{\rho_p}\right)^{1/5}. \label{eq:strength_v0} \end{equation} Since this estimate is based on the simple scaling analysis, only the rough order estimate is possible. In terms of the order estimate, the above estimate might be reasonable. However, there are some severe problems that must be overcome to fully understand the impact causing the wavy nano craters rim. Such problems are briefly discussed below. In general, the region of fingering instability depends on surrounding gas pressure~\cite{Xu} and wettability of the projectile~\cite{Duez} as well as impact inertia and properties of projectile and target. The droplet impact under vacuum condition cannot cause the fingering~\cite{Xu}. If the surface of projectile is hydrophilic, no-splashing is observed until the critical impact inertia~\cite{Duez}. These characteristics relate to the transition from the no-fingering to the clear-fingering states. While these effects might influence the value of parameters at the transition point, we consider the scaling exponent would not be affected. Furthermore, a recent work has revealed that the fingering of a droplet impact depends on the shape of the target~\cite{Paulo}. In the current investigation, we performed a very simple experiment and found that $N$ might be scaled by the combination of $We$, $Fr$ and $\rho_r$. This means that the crater deformation and the associated Rayleigh-Taylor instability are the possible sources of the fingering pattern formation. While only the viscous fluid targets are used in this study, viscoelastic or plastic behavior of the target must be also studied in order to discuss the general fingering instability. Moreover, the finite strength is indeed necessary to explain the observed result as mentioned above. Elastic property may work as a resistance to the splashing just like an atmospheric pressure which induces the fingering instability for the droplet impact~\cite{Xu}. Such solid-like elastic or plastic property is advantageous also to retain the crater shape on the target surface just like Itokawa's nano crater. However, much higher impact velocity and vacuum conditions are needed to study a viscoelastic or viscoplastic fingering. This is the important future problem. \section{Summary} We performed the low-velocity impact experiment with steel sphere projectiles and viscous target liquids. Milk-crown-like fingering can be observed in a particular parameter regime. The number of fingers created by the impact was counted from the high-speed video images. The observed fingers number $N$ can be scaled by the relation $N \simeq (\rho_r Fr)^{1/4}We^{1/2}/3^{3/4}$. This scaling is based on the Rayleigh-Taylor instability and the balance between the projectile's kinetic energy and the target cavity's potential energy. While the scaling reveals a certain aspect of the fingering induced by the impact to soft targets, further studies are necessary to clarify the details of impact fingering instability. Using a Rayleigh-Taylor-based scaling ($N \simeq D_c We^{1/2}/2\sqrt{3}D_p$), one can estimate the impact velocity for nano crater's wavy rim found on samples returned from the asteroid Itokawa. Assuming the strength-dominant cratering, the strength of the target material can also be estimated. However, more careful estimate by considering extreme conditions such as vacuum and low temperature is necessary to discuss the origin of wavy rim structures on Itokawa's nano craters. \section*{Acknowledgments} We would like to acknowledge S. Watanabe for introducing the Itokawa's nano crater shapes. This research has been partly supported by the JSPS, KAKENHI, number 23654134 and number 26610113.
\section{Introduction} A major area of current research is devoted to developing experimentally controllable systems that can be used for quantum computation, quantum communication, and quantum simulation of many-body physics. To date, most experiments have focused on the use of two-level systems (`qubits') for computation and communication \cite{Scarani2009,Ladd2010} and for the study of spin-1/2 (or spinless) many-body phenomena \cite{Bloch2012,Blatt2012}. However, there are a variety of motivations for performing experiments in higher-dimensional Hilbert spaces. Contrary to the intuition that enlarging the spin degree simplifies calculations by making them semiclassical \cite{Auerbach1994}, spin $> 1/2$ systems inherently have more complexity and cost exponentially more resources to classically simulate. For instance, it is computationally easy to find the ground state energy of a spin-1/2 chain with nearest-neighbor-only interactions in one-dimension; for systems with spin-7/2 or higher, the problem is known to belong to the QMA-complete complexity class, which is a quantum analogue of the classical NP-complete class \cite{Aharonov2007,HallgrenArxiv}. The difficulty of this problem for intermediate spin values, such as spin-1, is still an open question. From a more practical point of view, controllable three-level systems (`qutrits') are useful for quantum logic, since they can substantially simplify certain operations within quantum algorithms \cite{Lanyon2009} and can enhance the efficiency of quantum communication protocols \cite{Brukner2002}. When individual three-level systems are coupled together, they can be used to encode the physics of interacting \mbox{spin-1} particles. Such systems have attracted a great deal of theoretical interest following Haldane's conjecture that antiferromagnetic Heisenberg spin-1 chains, as opposed to spin-1/2 systems, have a finite energy gap that corresponds to exponentially decaying correlation functions \cite{Haldane1983,Haldane1983a}. This so-called Haldane phase possesses a doubly-degenerate entanglement spectrum \cite{Pollmann2010} and a non-local string order \cite{Kennedy1992,Kennedy1992a}, which is related to the order appearing in spin liquids \cite{Balents2010} and in the fractional quantum hall effect. These characteristics suggest that the Haldane phase is one of the simplest known examples of a symmetry-protected topological phase of matter \cite{Kennedy1992}. In addition to their interesting many-body properties, topological phases may be exploited in a more applied setting. The Haldane phase is useful for quantum operations (for instance, as a perfect quantum wire) \cite{Darmawan2010,Asoudeh2013} and can only be destroyed by crossing a phase transition. The finite energy gap in topological spin-1 systems makes them a potential candidate for long-lived, robust quantum memories \cite{Else2012NJP}, and schemes using symmetry-protected \mbox{spin-1} phases for measurement-based quantum computation have also been proposed \cite{Brennen2008,Else2012PRL}. Several groups have developed controllable three-level quantum systems by using pairs of photons \cite{Lanyon2008} or superconducting circuits \cite{Bianchetti2010} to implement qutrits, or by using spinor BECs to study quantum magnetism \cite{Chang2005,Stamper-Kurn2013,Parker2013}. However, no platform has yet used multiple interacting qutrits for quantum information protocols or for simulating lattice spin models. In this paper, we use trapped atomic ions to simulate a chain of \mbox{spin-1} particles with tunable, long-range XY interactions \cite{Cohen2014}. Our system performs the same basic tasks that are commonly used in spin-1/2 quantum simulations, such as observing dynamical state evolution \cite{Richerme2014}, measuring coherence and certifying entanglement \cite{Kim2010}, and adiabatically preparing nontrivial ground states \cite{Richerme2013}. With two spin-1 particles, we observe coherent evolution under the XY interactions among states in a `decoherence free' subspace \cite{Lidar1998,Kielpinski2001}. For certain states generated by the XY Hamiltonian, we can verify entanglement between a pair of 3-level systems with fidelities of up to 86\%. Adding a time dependent global field allows us to adiabatically prepare the ground state of the XY model for even numbers of spins. For odd numbers of spins, producing the calculated ground state is not possible with a simple adiabatic ramp since it requires crossing a first-order phase transition, hinting at the existence of a symmetry-protected phase. The tools demonstrated here could enable future studies of symmetry-protected order and can be extended to SU(3) models and other systems of higher symmetry \cite{Grass2013}. \FigureOne \section{Experimental implementation} The spin-1 chain is represented by a string of $^{171}$Yb$^+$ atoms held in a linear Paul trap. Three hyperfine levels in the $^2S_{1/2}$ ground manifold of each atom are used to encode the spin-1 states: $\ket{+} \equiv \ket{F=1,m_F=1}$, $\ket{-} \equiv \ket{F=1,m_F=-1}$, and $\ket{0} \equiv \ket{F=0,m_F=0}$, with frequency splittings of $\omega_\pm$ between the $\ket{0}$ and $\ket{\pm}$ states, as shown in Fig. \ref{fig:spin1levels}. Here, $\ket{+}$, $\ket{-}$, and $\ket{0}$ are the eigenstates of $S_z$ with eigenvalues +1, -1, and 0 respectively; $F$ and $m_F$ are quantum numbers associated with the total angular momentum of the atom and its projection along the quantization axis, defined by a magnetic field of $\sim$5 G. We apply global laser beams to the ion chain with a wavevector difference along a principal axis of transverse motion, driving stimulated Raman transitions between the $\ket{0}$ and $\ket{-}$ states and between the $\ket{0}$ and $\ket{+}$ states with equal Rabi frequencies $\Omega_i$ on ion $i$ \cite{Kim2009}. To generate spin-1 XY interactions, we apply two beat frequencies at $\omega_- + \mu$ and $\omega_+ - \mu$ to these respective transitions, where $\mu$ is slightly detuned from the transverse motional frequencies, as shown in Fig. 1(c). Under the rotating wave approximations $\omega_\pm \gg \mu \gg \Omega_i$ and within the Lamb-Dicke regime ($\Delta k \left< \hat{x}_i \right> \ll 1$, with $\Delta k$ the wavevector difference of the Raman beams and $\hat{x}_i$ the position operator of the $i$th ion), the resulting interaction Hamiltonian (with $h=1$) is \eq H = \sum_{i,m=1}^N \frac{i\eta_{i,m}\Omega_i}{2\sqrt{2}} \left( -S^i_+ a_m e^{i(\mu-\omega_m) t} + S^i_- \adag_m e^{-i(\mu-\omega_m) t} \right). \eeq Here $a_m$ and $\adag_m$ are the phonon operators of the normal mode $m$ with frequency $\omega_m$, $\eta_{i,m} = b_{i,m}\sqrt{\hbar (\Delta k)^2/2 M \omega_m}$ is the Lamb-Dicke factor (where $b_{i,m}$ is the normal mode transformation matrix \cite{James1998} and $M$ is the mass of a single ion), and the spin raising and lowering operators $S^i_\pm$ satisfy the commutation relations $\left[S_+^i, S_-^j\right] = 2 S_z^i \delta_{ij}$. In the limit where the beatnotes are far detuned ($\eta_{i,m} \Omega_i \ll |\mu - \omega_m|$) and the phonons are only virtually excited, this results in an effective Hamiltonian with XY-type spin-spin interactions and spin-phonon couplings, \begin{eqnarray} \nonumber H_{\mathrm{eff}} = &\sum_{i<j}& \frac{J_{i,j}}{4} \left(S_+^i S_-^j + S_-^i S_+^j \right) \\ \label{eqn:XYHam} + &\sum_{i,m}& V_{i,m} \left[ \left(2 a_m^\dagger a_m +1 \right) S_z^i - \left( S_z^i \right)^2 \right]. \label{eq:spin1XY} \end{eqnarray} The pure spin-spin interaction in the first term of Eq. \ref{eq:spin1XY} follows the same formula as for generating spin-1/2 Ising interactions \cite{Kim2009}: \eq J_{i,j} = \Omega_i \Omega_j \sum_m \frac{\eta_{i,m} \eta_{j,m}}{2(\mu-\omega_m)}. \label{eq:Jij} \eeq When $\mu$ is larger than the transverse center-of-mass frequency, $J_{i,j}$ falls off with distance as roughly $J_{i,j} \sim J_0/|i-j|^\alpha$, where $J_0$ is of order $\approx 1$ kHz and $\alpha$ can be tuned between 0 and 3 using trap and laser parameters \cite{Porras2004,Islam2013}. The $V_{i,m}$ term in Eq. \ref{eqn:XYHam} is given by a similar formula, \eq V_{i,m} = \frac{\left(\eta_{i,m} \Omega_i\right)^2}{8(\mu-\omega_m)}. \eeq For very long-ranged spin-spin interactions ($\alpha \lesssim 0.5$), or for small numbers of ions, the $V_{i,m}$ terms are approximately uniform across the spin chain. In these instances, the $V_{i,m}$ coefficient can be factored out of the sum over ions in Eq. \ref{eq:spin1XY}, leaving only global $S_z^i$ and $(S_z^i)^2$ terms. For shorter-range interactions or for longer chain lengths, the $V_{i,m}$ terms can be eliminated by adding an additional set of beat frequencies at $\omega_- - \mu$ and $\omega_+ + \mu$, which would generate Ising-type interactions between effective spin-1 particles using the M$\o$lmer-S$\o$rensen gate \cite{Molmer1999}. \FigureTwo The ions are initialized before each experiment by cooling the transverse modes near their ground state of motion ($\bar{n}\approx 0.05$) and optically pumping the spins to the $\ket{00\cdots}$ state. After applying the Hamiltonian in Eq. \ref{eq:spin1XY} for varying lengths of time, we measure the population of the state $\ket{0}$ at each site by imaging spin-dependent fluorescence \cite{Olmschenk2007} onto an intensified CCD camera and observing which ions are `dark'. Because both of the $\ket{\pm}$ states appear `bright' during the detection process and are scattered into an incoherent mixture of the $\ket{F=1}$ states, our current setup does not allow discrimination among all three possible spin states in a single experiment. However, we can measure the population of either $\ket{+}$ or $\ket{-}$ by repeating the experiment and applying a $\pi$ rotation to the appropriate $\ket{0}\leftrightarrow\ket{\pm}$ transition before the fluorescence imaging. For instance, measuring an ion in the `dark' state after a $\pi$ pulse between $\ket{0}\leftrightarrow\ket{+}$ indicates that the spin was in the $\ket{+}$ state before detection. This binary discrimination is not a fundamental limit to future experiments, since populations could be `shelved' into atomic states that do not participate in the detection cycle. Since the ions are initialized to the $\ket{00\cdots}$ state, and because the spin-spin interactions in Eq. \ref{eq:spin1XY} conserve the quantity $\sum_i S_z^i \equiv \mathcal{S}_z$, the dynamics are restricted to the set of states with $\szsubspace$. The $\szsubspace$ subspace is protected against fluctuations in the real magnetic field $\Delta B(t)$, which would otherwise result in an unwanted noise term $\mu_B \Delta B(t) \mathcal{S}_z$ (where $\mu_B$ is the Bohr magneton). For instance, the $T_2$ coherence times of the $\ket{0}\leftrightarrow\ket{\pm}$ transitions were measured to be 0.5 ms, limited by magnetic field noise. Nevertheless, the data in Figures \ref{fig:dynamics2ions} and \ref{fig:ParityFig} (below) exhibit coherence and entanglement for several ms (limited by laser intensity noise), demonstrating the robustness of this `decoherence-free' subspace against time-varying magnetic fields. Remaining within this subspace does not substantially limit the size of the accessible Hilbert space, since the number of states in the $\szsubspace$ subspace of $N$ spin-1 particles scales as $\sim 3^N/(2\sqrt{N})$ for large $N$, which is exponentially greater than the $2^N$ states accessible in a spin-1/2 system. \section{Coherent dynamics of two spins and entanglement verification} For a system of 2 spins, dynamical evolution under the Hamiltonian in Eq. \ref{eq:spin1XY} can be understood as Rabi flopping between the $\ket{00}$ and $\left( \ket{+-} + \ket{-+} \right)/\sqrt{2}$ states with Rabi frequency $\sqrt{2} J_{1,2}$. This behavior is shown in Fig. \ref{fig:dynamics2ions}, where panels (a), (b), and (c) show the probability of each ion to be in the $\ket{0}$, $\ket{-}$, and $\ket{+}$ states, respectively. The population remains in the $\mathcal{S}_z=0$ subspace, as expected: Fig. \ref{fig:dynamics2ions}(a) shows the absence of the $\mathcal{S}_z \neq 0$ states ($\ket{0+}$, $\ket{0-}$, $\ket{+0}$, and, $\ket{-0}$), while Fig. \ref{fig:dynamics2ions}(b) and (c) respectively show the absence of the other $\mathcal{S}_z \neq 0$ states $\ket{--}$ and $\ket{++}$. The drift in $J_{1,2}$ evidenced in Fig. \ref{fig:dynamics2ions}(c) could be stabilized in future experiments by feeding back to the trap RF voltage to better stabilize the radial trap frequencies. The different ions $i$ can experience position-dependent $S_z^i$ and $\left(S_z^i\right)^2$ shifts. We attribute this effect to a micromotion gradient, since the shifts can be compensated by adjustments of the voltages on the DC trap electrodes. The calculation overlaid in Fig. \ref{fig:dynamics2ions} includes the site-dependent terms $(200 $ Hz$) S_z^{(2)}$ + $(150$ Hz$) (S_z^{(2)})^2$, which were left as free fitting parameters when numerically evolving the Schr\"odinger equation under the Hamiltonian in Eq. \ref{eqn:XYHam}. The plotted curves assume strictly unitary evolution (i.e. no decoherence) over the timescale of the experiments. \FigureThree At a time $t=0.5/(\sqrt{2}J_{1,2})$, which is roughly 0.27 ms in Fig. \ref{fig:dynamics2ions}, the system is left approximately in the entangled state $\left( \ket{+-} + \ket{-+} \right)/\sqrt{2}$. To verify entanglement in the system, one could use spin-1 analogues of Bell-type inequalities \cite{Collins2002}, which require many local rotations but are sensitive to maximally entangled states like $\left(\ket{00}+\ket{+-}+\ket{-+}\right)/\sqrt{3}$. However, for the class of states generated by the XY interactions, a much simpler series of global rotations is sufficient to verify entanglement. The analysis consists of performing three sequential rotations on the $\ket{0}$ to $\ket{\pm}$ transitions, \eq \label{eqn:Rotations} R_{0\pm}(\theta,\phi) = e ^{\left( \frac{i\theta}{2} \sum_k [e^{\pm i\phi} (\ket{\pm}\!\bra{0})_k + e^{\mp i\phi} (\ket{0}\!\bra{\pm})_k] \right)}, \eeq before measuring the population in $\ket{0}$. The rotation sequence is given by $R_{0+}(\pi/2,\varphi)R_{0+}(\pi/2,0)R_{0-}(\pi,0)$, with the rotations applied from right to left. The first two rotations map the state $\left( \ket{+-} + \ket{-+} \right)/\sqrt{2}$ to $\left( \ket{00} + \ket{++} \right)/\sqrt{2}$, while the phase of the third rotation is varied to analyze the entanglement of this resulting state \cite{Sackett2000}. The parity $\Pi = \sum_{j=0}^2 (-1)^j P_j$ (with $P_j$ the probability of $j$ atoms in $\ket{0}$) oscillates as a function of the phase $\varphi$ of the third pulse, and the amplitude of its oscillation depends on the off-diagonal density matrix elements: \small \begin{eqnarray} \Pi(\varphi) =& C + \frac{1}{2}\cos 2\varphi \left( P_{\mathsmaller{--}}+ P_{\mathsmaller{++}} - P_{\mathsmaller{+-}} - P_{\mathsmaller{-+}} \right. \\ &\left. - 2 |\rho_\mathsmaller{+-,-+}| - 2 |\rho_\mathsmaller{--,++}| \right) \nonumber \\ + \frac{1}{2} \sin& 2\varphi \left( 2|\rho_\mathsmaller{-+,++}| + 2|\rho_\mathsmaller{+-,++}| - 2|\rho_\mathsmaller{--,-+}| - 2|\rho_\mathsmaller{-+,--}| \right), \nonumber \end{eqnarray} \normalsize where $P_{\mathsmaller{i}}$ is the population in state $\ket{i}$ ($\ket{i} = \ket{--},\ket{-+}$, etc.), $\rho_{i,j}$ is the off-diagonal density matrix element quantifying the coherence between $\ket{i}$ and $\ket{j}$, and $C$ is a constant offset that depends on the various density matrix elements but not on the phase $\varphi$ of the final rotation. The populations in $\ket{++}$ and $\ket{--}$ are negligible, simplifying this expression: \begin{equation} \Pi(\varphi) \approx C - A\cos2\varphi \end{equation} where the oscillation amplitude \eq A = \frac{1}{2} \left(P_{\mathsmaller{+-}} + P_{\mathsmaller{-+}} + 2 |\rho_\mathsmaller{+-,-+}| \right) \eeq is akin to the entanglement fidelity $\mathcal{F}$ of GHZ states in two-level systems \cite{Sackett2000}. Measuring the amplitude $A$ of the parity oscillation $\Pi(\varphi)$ then allows us to verify entanglement for certain classes of states. According to an analysis analogous to that in \cite{Sackett2000}, the following inequality holds for all separable qutrit states: \eq 2A + P_{\mathsmaller{00}} + 2 |\rho_{\mathsmaller{+-,00}}| + 2 |\rho_{\mathsmaller{-+,00}}| \leq 1. \label{eq:entanglementInequality} \eeq Hence, violation of this inequality demonstrates entanglement between spin-1 particles or qutrits, and measuring an amplitude of $A> 1/2$ is sufficient to violate the inequality. Figure \ref{fig:ParityFig}(b) shows an example of the measured parity curve used to extract the amplitude $A$ and verify entanglement between the qutrit pair. Such measurements can be repeated for different durations of exposure to the XY Hamiltonian. At times $t = (2n+1)/(2\sqrt{2}J_{1,2})$ ($n=0,1,2,\ldots$), the system should again be in the state $\left( \ket{+-} + \ket{-+} \right)/\sqrt{2}$, while at times $t=n/(\sqrt{2}J_{1,2})$ it should return to the unentangled product state $\ket{00}$. The result is plotted in Fig. \ref{fig:ParityFig}(c). Two known sources of dephasing contribute to the observed loss of coherence in the experiment. First, laser intensity fluctuations and pointing instability cause noise in the spin-spin coupling term, leading to apparent dephasing when many repetitions are averaged together. These fluctuations could be compensated in future experiments by variants of the method of composite pulses \cite{Brown2004, Albrecht2014}. The second dephasing source results from inhomogeneities in the $V_{i,m}$ term (Eq. \ref{eqn:XYHam}) across the chain, which will cause different spins to acquire phases at different rates. This could be compensated by adding an extra driving term to cancel the inhomogeneities or by applying a series of echo pulses \cite{Cai2012,Biercuk2009PRA}. Fluctuating external magnetic fields and off-resonant coupling to the carrier transition would ordinarily add dephasing noise along the $\hat{z}$ direction, but have been suppressed here by working in the $\mathcal{S}_z=0$ subspace. \FigureFour \FigureFive \section{Ground state production} We can also add an effective $(S_z^i)^2$ field term, $D \sum_{i=1}^N (S_z^i)^2$, to the Hamiltonian by shifting the beat frequencies of the Raman lasers to $\omega_+ - \mu - D$ and $\omega_- + \mu - D$. This $(S_z^i)^2$ term can be used to adiabatically prepare the ground state of the XY Hamiltonian in Eq. \ref{eq:spin1XY}. As before, the spins are prepared in $\ket{00\cdots}$, which is the approximate ground state of Eq. \ref{eqn:XYHam} in the presence of a large (5 kHz) $(S_z^i)^2$ field. This field is then ramped down slowly according to $D(t) = (5 \:\mathrm{kHz}) e^{-t/(0.167 \:\mathrm{ms})}$. Figure \ref{fig:EvenIonGroundState} shows the populations measured at the end of the $(S_z^i)^2$ ramp for two and four spins, which match reasonably well with the calculated ground state. Measurements of populations in the $S_z$ basis necessarily discard phase information about components of the final state. This can be important in many spin models, including the XY model, where such measurements alone cannot discriminate between different eigenstates. For example, the ground state of an XY model with two spin-1 particles is $\ket{00}/\sqrt{2} - \left( \ket{-+} + \ket{+-} \right)/2$, while the highest excited state is $\ket{00}/\sqrt{2} + \left( \ket{-+} + \ket{+-} \right)/2$, differing only by a relative phase. We check that we are creating the ground state after our adiabatic protocol by applying a pair of rotations, $R_{0-}(\pi/2,\varphi)R_{0+}(\pi/2,0)$, and measuring the parity $\Pi$ as was done in the entanglement analysis. This is expected to result in $\Pi(\varphi) = \frac{3}{8} \pm \frac{1}{2}\cos\varphi$, where the + and - correspond to the ground and highest excited states, respectively. As shown in Fig. \ref{fig:GroundStatePhase}, our measurements are consistent with having prepared the 2-spin ground state. \section{Toward Haldane physics} A more long-term goal for spin-1 quantum simulations will be to produce and study ground states in the Haldane phase \cite{Cohen2014}. It is known that an XY model with both nearest-neighbor and next-nearest-neighbor interactions can exhibit a symmetry-protected Haldane phase \cite{Murashima2005}, and it remains an open question whether a generic long-range XY model would show the same behavior. Already with our experimentally implemented Hamiltonian, we find a useful test case where the symmetry of the ground state prevents it from being created via the simple adiabatic protocol described above. The ground state $\ket{\psi}_{gs}$ of a long-range XY model can be calculated exactly for three spins. For our experimental coupling strengths $J_{i,j} \sim 1/|i-j|^{0.36}$, \eqarray \nonumber \ket{\psi}_{gs} = &\sqrt{0.16}& \left(\ket{0-+} -\ket{0+-} + \ket{-+0} - \ket{+-0} \right) \\ +& \sqrt{0.18}& \left(\ket{+0-} - \ket{-0+}\right). \label{eqn:XYgroundstate} \eeqarray This state has a 99.9\% overlap with a three-spin AKLT state \cite{Affleck1988}, which is the canonical example of a ground state in the Haldane phase that can be written in closed form for any number of spins. The state in Eq. \ref{eqn:XYgroundstate} is antisymmetric with respect to the same symmetries that govern the Haldane phase, such as left-right spatial inversion of the chain or a global rotation about $S_x$ by $\pi$ (which sends $\ket{+}$ to $\ket{-}$ and vice versa). However, since the starting state $\ket{000}$ and the applied Hamiltonian are symmetric with respect to these operations, we should be unable to reach the antisymmetric ground state with a simple adiabatic ramp. Indeed, we find numerically that a first order phase transition separates the symmetric and antisymmetric ground states, which cannot be adiabatically connected without breaking inversion and rotational symmetry. For the three-spin experiment in Fig. \ref{fig:3ionGroundState}, we hence prepare a state close to the first excited state rather than the ground state. This observation suggests that even in the presence of various experimental imperfections, the ground state of our three-spin XY model enjoys the same symmetry protection as the Haldane phase. \FigureSix In this paper, we have demonstrated the basic ingredients which are needed for the implementation of quantum simulations with spins greater than 1/2. We believe that this work opens paths for studying the exciting physics beyond spin-1/2 systems, and we have already taken the first steps towards exploring the richness of topological phases. In particular, for a long-range spin-1 XY model, we have demonstrated coherent Schr\"odinger evolution and the capability to create symmetric ground states. We have observed that for odd numbers of spins, symmetry considerations prevent us from creating ground states which bear a close resemblance to AKLT states and hence may belong to the Haldane phase. Future work will address the questions of how to add a Heisenberg term and symmetry-breaking perturbations to the Hamiltonian so as to prepare antisymmetric ground states \cite{Cohen2014}, which will allow us to create and probe interesting edge states in the Haldane phase. \section{Acknowledgments} We thank Brian Neyenhuis, Paul Hess, Alexey Gorshkov, and Zhe-Xuan Gong for critical discussions. This work is supported by the U.S. Army Research Office (ARO) Award W911NF0710576 with funds from the DARPA Optical Lattice Emulator Program, ARO award W911NF0410234 with funds from the IARPA MQCO Program, and the NSF Physics Frontier Center at JQI.
\section{Introduction} \label{intro} The number of TeV blazars has grown rapidly in recent years (see, e.g., the reviews of Holder 2012, 2014) with 54 TeV blazars currently known\footnote{\url{http://tevcat.uchicago.edu/}}. The vast majority of these (44 of 54, or about 80\%) belong to the blazar sub-class of high-frequency peaked BL Lac objects, or HBLs. Several of these TeV HBLs have displayed remarkable variability in their TeV gamma-ray emission on time scales as short as a few minutes (e.g., Aharonian et al.\ 2007; Albert et al.\ 2007; Sakamoto et al.\ 2008). Although various explanations have been proposed for such rapid variability (e.g., Begelman et al.\ 2008; Nalewajko et al.\ 2011; Narayan \& Piran 2012; Barkov et al.\ 2012), they share the common feature of high bulk Lorentz factors of at least $\gtrsim$25 for the gamma-ray emitting plasma in their relativistic jets. High bulk Lorentz factors and Doppler factors are also required to model TeV blazar spectral energy distributions (e.g., Tavecchio et al.\ 2010), particularly in the case of one-zone models. For example, fitting the SED of the TeV blazar PKS~1424+240 with a one-zone model yields a Doppler factor of $\delta\sim100$ (Aleksi{\'c}, et al.\ 2014a). Direct imaging of the jets of these blazars on parsec-scales requires VLBI. Most HBLs are relatively faint in the radio, so the TeV HBLs are not well represented in large VLBI monitoring projects, such as MOJAVE (Lister et al.\ 2009). We have previously reported multi-epoch VLBI kinematic results for 11 established TeV HBLs (Piner et al.\ 2010; Tiet et al.\ 2012). A major result of those kinematic analyses was the absence of any rapidly moving features in the jets of those blazars; all components in all 11 sources were either stationary or slowly moving ($\lesssim$1$c$). Slow apparent speeds of VLBI components in specific TeV HBLs has been confirmed by numerous other studies (e.g., Giroletti et al.\ 2004a; Lico et al.\ 2012; Blasi et al.\ 2013; Aleksi{\'c} et al.\ 2013; Richards et al.\ 2013), although note that TeV-detected intermediate-peaked BL Lac objects (IBLs), such as 3C 66A and BL Lac, do show apparently superluminal components (e.g., Britzen et al.\ 2008). While effects other than slow bulk motion can produce slow apparent speeds of components, the complete absence of any rapidly moving features in all of these jets, after as much as 20 years of VLBA monitoring (for Mrk 421 and Mrk 501), and even after powerful flares (Richards et al.\ 2013), is quite distinct from the behavior of other types of gamma-ray blazars, which show frequent superluminal ejections (e.g., Lister et al.\ 2009; Marscher 2013). Taken together with other measured radio properties, such as the brightness temperatures and core dominance (Giroletti et al.\ 2004b; Lister et al.\ 2011), the VLBI data imply only modest bulk Lorentz factors and Doppler factors in the parsec-scale radio jets of these TeV HBLs. (Note that, because the sources appear one-sided on parsec scales, the VLBI data do require that the sources be at least moderately relativistic.) This discrepancy between the Doppler and Lorentz factors estimated from the gamma-ray data and the radio date has been referred to as the ``Doppler Crisis'' of TeV blazars. A natural explanation for the Doppler Crisis is that the radio and gamma-ray emission may be produced in different parts of the jet with different bulk Lorentz factors. Several variations of such a multi-component jet have been proposed including decelerating jets (Georganopoulos \& Kazanas 2003), spine-sheath structures (Ghisellini et al.\ 2005), `minijets' within the main jet (Giannios et al.\ 2009), and faster moving leading edges of blobs (Lyutikov \& Lister 2010), but they all require that the jets of HBLs contain significant velocity structures. Some of these velocity structures, such as a fast spine and slower layer, may under certain conditions produce observable signatures in VLBI images, such as limb brightening of the transverse jet structure. Limb brightening has indeed been observed in VLBI images of the bright TeV blazars Mrk 421 and Mrk 501 (e.g., Giroletti et al.\ 2004a, 2006, 2008; Piner et al.\ 2009, 2010; Croke et al.\ 2010; Blasi et al.\ 2013). These arguments for velocity structures in the jets of TeV HBLs are independently supported by developments in radio-loud AGN unification (Meyer et al.\ 2011, 2013a; see also Ghisellini et al.\ 2009). In that unification work, radio-loud AGN are divided into two distinct sub-populations that constitute a `broken power sequence'. The `weak' jet sub-population resulting from inefficient accretion modes (and corresponding to HBLs when viewed at a small angle) follows a de-beaming curve that requires velocity gradients in the jets, such as a decelerating or spine-sheath jet; see also the similar arguments in earlier unification work (Chiaberge et al.\ 2000). The TeV HBLs may thus represent the small viewing angle peak of a distinct radio-loud population with both fundamentally different jet structure and accretion mode from the more powerful blazars. If this is the case, then obtaining more information on the parsec-scale structure of these sources through high-resolution imaging is quite important. We are presently taking advantage of both the rapidly growing TeV blazar source list and the recently upgraded sensitivity of the Very Long Baseline Array (VLBA) to significantly expand our previous work on the parsec-scale structure of TeV HBLs (e.g., Piner et al.\ 2010; Tiet et al.\ 2012). Here, we present first-epoch VLBA images of twenty newer TeV HBLs discovered during the years 2006 to 2013, several of which had never been imaged with VLBI. This represents the first stage of a multi-epoch VLBA monitoring program on these sources designed to provide parsec-scale kinematic and structural information on nearly the complete sample of TeV HBLs. In $\S$~\ref{observations} we describe the source selection and observations, in $\S$~\ref{results20} we present the results for these observations, and in $\S$~\ref{resultsall} we extend consideration to the full set of TeV HBLs. Final discussion and conclusions are given in $\S$~\ref{discussion}. \section{Observations} \label{observations} \subsection{Source Selection} \label{sourceselection} We have been conducting VLBA observations of TeV-detected HBLs since the discovery of the first two TeV blazars (Mrk~421 and Mrk~501) in the 1990s, in order to study their jet physics through high-resolution parsec-scale imaging (see $\S$~\ref{intro}). Our complete candidate source list is thus the 44 HBLs listed as detections in the TeVCat catalog\footnote{\url{http://tevcat.uchicago.edu/}} as of this writing. From those 44 sources we excluded the following for the observations in this paper: \begin{enumerate} \item{Eleven sources reported as TeV detections before 2007 for which we have already published multi-epoch VLBA observations: six of these sources are discussed by Piner et al.\ (2010), and an additional five by Tiet et al.\ (2012).} \vspace*{-0.05in} \item{Seven sources with sufficient multi-epoch VLBA data in the MOJAVE monitoring program \footnote{\url{http://www.physics.purdue.edu/astro/MOJAVE/\\allsources.html}}.} \vspace{-0.05in} \item{Three sources which are below $-40\arcdeg$ declination, and thus difficult to image with the VLBA.} \vspace{-0.05in} \item{Two sources which were detected too recently (after 2013) to be included in this work.} \vspace{-0.05in} \item{The low brightness temperature source HESS~J1943+213 (Gab{\'a}nyi et al.\ 2013).} \end{enumerate} These exclusions are shown in tabular from in Table~\ref{selecttab}, and they leave 20 HBLs (or nearly half of the full sample) that were all reported as new detections by the TeV telescopes between 2006 and 2013, and that have not yet been studied with multi-epoch VLBI imaging by any program. The goal of the observations presented here is to provide high-dynamic range single-epoch images of these 20 sources, as a precursor to a multi-epoch monitoring program to study the jet kinematics. \begin{table*}[!t] \begin{center} {\normalsize \caption{Sample Selection} \label{selecttab} \begin{tabular}{l c c l c c} \tableline \tableline \\[-5pt] \multicolumn{1}{c}{Source$^{a}$} & Included$^{b}$ & Reason$^{c}$ & \multicolumn{1}{c}{Source$^{a}$} & Included$^{b}$ & Reason$^{c}$ \\ \tableline \\[-5pt] SHBL~J001355.9$-$185406 & Y & ... & Markarian~421 & N & 1 \\ [2pt] KUV~00311$-$1938 & Y & ... & Markarian~180 & N & 1 \\ [2pt] 1ES~0033+595 & Y & ... & RX~J1136.5+6737 & N & 4 \\ [2pt] RGB~J0136+391 & Y & ... & 1ES~1215+303 & N & 2 \\ [2pt] RGB~J0152+017 & Y & ... & 1ES~1218+304 & N & 1 \\ [2pt] 1ES~0229+200 & Y & ... & MS~1221.8+2452 & Y & ... \\ [2pt] PKS~0301$-$243 & N & 2 & 1ES~1312$-$423 & N & 3 \\ [2pt] IC~310 & N & 2 & PKS~1424+240 & N & 2 \\ [2pt] RBS~0413 & Y & ... & H~1426+428 & N & 1 \\ [2pt] 1ES~0347$-$121 & Y & ... & 1ES~1440+122 & Y & ... \\ [2pt] 1ES~0414+009 & Y & ... & PG~1553+113 & N & 1 \\ [2pt] PKS~0447$-$439 & N & 3 & Markarian~501 & N & 1 \\ [2pt] 1ES~0502+675 & Y & ... & H~1722+119 & Y & ... \\ [2pt] PKS~0548$-$322 & Y & ... & 1ES~1727+502 & N & 2 \\ [2pt] RX~J0648.7+1516 & Y & ... & 1ES~1741+196 & Y & ... \\ [2pt] 1ES~0647+250 & Y & ... & HESS~J1943+213 & N & 5 \\ [2pt] RGB~J0710+591 & Y & ... & 1ES~1959+650 & N & 1 \\ [2pt] 1ES~0806+524 & N & 2 & PKS~2005$-$489 & N & 3 \\ [2pt] RBS~0723 & N & 4 & PKS~2155$-$304 & N & 1 \\ [2pt] 1RXS~J101015.9$-$311909 & Y & ... & B3~2247+381 & Y & ... \\ [2pt] 1ES~1011+496 & N & 2 & 1ES~2344+514 & N & 1 \\ [2pt] 1ES~1101$-$232 & N & 1 & H~2356$-$309 & N & 1 \\ \tableline \\[-10pt] \end{tabular}} \end{center} {\bf Notes.}\\ $a$: Source names are the so-called `Canonical Name' used by TeVCat.\\ $b$: Whether or not the source is included in the new observations for this paper.\\ $c$: Reason for exclusion: 1:~Monitored in our previous work, 2:~in MOJAVE program, 3:~too far south, 4:~detection too recent, 5:~low brightness temperature (see $\S$~\ref{sandtb}).\\ \end{table*} Single-epoch pilot images of 8 of these 20 sources, obtained prior to the VLBA sensitivity upgrade (Romney et al.\ 2009), were presented by Piner \& Edwards (2013). For those 8 sources, we present in this paper the first epoch of a multi-epoch monitoring series obtained after the sensitivity upgrade, and with images of significantly higher dynamic range compared to those in Piner \& Edwards (2013). For the remaining 12 sources, we present the pilot images made to assess suitability for multi-epoch monitoring, all of which were made subsequent to the VLBA sensitivity upgrade. The VLBI and gamma-ray properties of the entire sample of 44 TeV HBLs are discussed later in this paper (see $\S$~\ref{resultsall} and Table~\ref{alltab}). \subsection{Details of Observations} \label{obsdetails} Details of the observing sessions are given in Table~\ref{obstab}. All observations were made at an observing frequency of 8.4~GHz (4~cm), because this provides the optimum combination of angular resolution and sensitivity for these fainter sources. All observations used the full 2 Gbps recording rate of the VLBA, and were made using the polyphase filterbank (PFB) observing system of the Roach Digital Backend (RDBE), in its dual-polarization configuration of eight contiguous 32~MHz channels at matching frequencies in each polarization. Although dual-polarization was recorded, only total intensity (Stokes I) was calibrated and imaged, because of the likely sub-millijansky level of polarized flux from most of these sources. We used phase-referencing for three of the fainter targets: SHBL~J001355.9$-$185406, 1ES~0347$-$121, and 1RXS~J101015.9$-$311909; both because their correlated flux densities were uncertain, and to obtain precise milliarcsecond-scale positions because they were not in the VLBA input catalog. These data were phase-referenced to the ICRF sources J0015$-$1812, J0351$-$1153, and J1011$-$2847 respectively, all of which had separations of less than $3\arcdeg$ from the target sources. Based on the ICRF positions of the calibrator sources, we derived (J2000) positions of R.A.~=~00$^{\rm {h}}$13$^{\rm {m}}$56.043$^{\rm {s}}$, decl.~=~$-$18$\arcdeg$54$'$06.696$''$ for SHBL~J001355.9$-$185406, R.A.~=~03$^{\rm {h}}$49$^{\rm {m}}$23.186$^{\rm {s}}$, decl.~=~$-$11$\arcdeg$59$'$27.361$''$ for 1ES~0347$-$121, and R.A.~=~10$^{\rm {h}}$10$^{\rm {m}}$15.979$^{\rm {s}}$, decl.~=~$-$31$\arcdeg$19$'$08.408$''$ for SHBL~J001355.9$-$185406, which we expect to be accurate to a few milliarcseconds. Although those observations were done in phase-referencing mode, all three sources were bright enough for fringe-fitting, and the fringe-fit data were used in the subsequent imaging. \begin{table*}[!t] \begin{center} {\normalsize \caption{Observation Log} \label{obstab} \begin{tabular}{l c c c c} \tableline \tableline \\[-5pt] \multicolumn{1}{c}{Date} & Observation & Observing & Excluded & \multicolumn{1}{c}{Target Sources} \\ & Code & Time & VLBA & \\ & & (hours) & Antennas$^{a}$ & \\ \tableline \\[-5pt] 2013 Aug 16 & S6117D1 & 6 & FD,LA & SHBL~J001355.9$-$185406, 1ES~0033+595 \\ [3pt] 2013 Aug 23 & S6117A1 & 8 & None & RGB~J0152+017, 1ES~0229+200, \\ [2pt] & & & & RBS~0413, 1ES~0347$-$121 \\ [3pt] 2013 Aug 30 & S6117D2 & 6 & None & KUV~00311$-$1938, RGB~J0136+391\\ [3pt] 2013 Sep 19 & S6117B1 & 8 & KP & 1ES~0414+009, 1ES~0502+675, \\ [2pt] & & & & PKS~0548$-$322, RGB~J0710+591 \\ [3pt] 2013 Oct 21 & S6117D3 & 6 & LA & RX~J0648.7+1516, 1ES~0647+250 \\ [3pt] 2013 Oct 24 & S6117D4 & 6 & FD,LA & 1RXS~J101015.9$-$311909, MS~1221.8+2452 \\ [3pt] 2013 Dec 23 & S6117D5 & 9 & KP,NL & 1ES~1440+122, H1722+119, \\ [2pt] & & & & 1ES~1741+196, B3~2247+381 \\ \tableline \\[-10pt] \end{tabular}} \end{center} {\bf Notes.}\\ $a$: VLBA antennas that did not participate or that were excluded from the imaging for that session. FD=Fort Davis, Texas, KP=Kitt Peak, Arizona, LA=Los Alamos, New Mexico, NL=North Liberty, Iowa.\\ \end{table*} We used the AIPS software package for calibration and fringe-fitting of the correlated visibilities, and fringes were found across the full bandwidth at significant SNR and small delays and rates to all target sources. A small number of discrepant visibilities were flagged, and the final images were produced using CLEAN and self-calibration in the DIFMAP software package. VLBA imaging of sources at these lower flux density levels can be sensitive to the self-calibration averaging interval, and self-calibration will generate spurious point-source structure if the averaging interval is too short (e.g., Mart{\'{\i}}-Vidal \& Marcaide 2008). We carefully investigated and selected self-calibration solution intervals for the fainter sources to make sure that minimal spurious flux density (less than $\sim1$~mJy) should be introduced into the images through self-calibration (see Equations~7 and 8 of Mart{\'{\i}}-Vidal \& Marcaide 2008). In the section below, all images are displayed using natural weighting, in order to maximize the dynamic range. At a typical redshift for these sources of $z\sim0.2$, 1~milliarcsecond (a typical beam size) corresponds to a linear resolution of about 3~parsecs, and the smallest sizes measurable in model fitting (about 10\% of the beam size) would have a linear size of about 0.3~parsecs. \section{Results for the 20 New Sources} \label{results20} \subsection{Images} \label{images} \begin{figure*}[!t] \centering \includegraphics[scale=0.85]{f1_p1.ps} \caption{VLBA images at 8.4~GHz of TeV blazars from Table~\ref{imtab}. Parameters of the images are given in Table~\ref{imtab}. Axes are in milliarcseconds. The lowest contour in each images is 3 times the rms noise level from Table~\ref{imtab}, and each subsequent contour is a factor of two higher.} \end{figure*} \begin{figure*}[!t] \centering \includegraphics[scale=0.85]{f1_p2.ps} \vspace*{0.125in} \\Fig. 1.--—{\em Continued} \end{figure*} The VLBA images of the 20 TeV HBLs studied for this paper are shown in Figure~1, and the parameters of these images are tabulated in Table~\ref{imtab}. The B1950 name is shown in each panel in Figure~1, and may be used subsequently to refer to the source. All sources show a bright, compact component, hereafter identified as the VLBI core, and they all show additional extended structure that can be modeled by at least one Gaussian feature in addition to the core (see $\S$~\ref{mfits}). Thus, all of these sources are suitable for continued VLBI monitoring to study the parsec-scale jet kinematics. The images in Figure~1 do not show the entire CLEANed region for clarity, but instead are zoomed in on the core and the inner jet region. Larger scale images plus all associated data files are available at the project web site \footnote{www.whittier.edu/facultypages/gpiner/research/archive/\\archive.html}. The peak flux densities in the images in Figure~1 range from 7 to 98 mJy~bm$^{-1}$ (see Table~\ref{imtab}). However, the noise levels are quite low, typically only about 0.02 mJy~bm$^{-1}$ (Table~\ref{imtab}), close to the expected thermal noise limit for these observations, so that even the images of the fainter sources have dynamic ranges of several hundred, which is easily high enough to image the parsec-scale jet structure. \begin{table*}[!t] \begin{center} {\normalsize \caption{Parameters of the Images} \label{imtab} \begin{tabular}{l c c c c c} \tableline \tableline \\ [-5pt] \multicolumn{1}{c}{Source} & B1950 & Time On & Beam & Peak Flux & $I_{\rm {rms}}^{b}$ \\ & Name & Source & Parameters$^{a}$ & Density & (mJy bm$^{-1}$) \\ & & (minutes) & & (mJy bm$^{-1}$) \\ \tableline \\ [-5pt] SHBL~J001355.9$-$185406 & 0011$-$191 & 120 & 2.15,0.82,$-$4.1 & 10 & 0.029 \\ [2pt] KUV~00311$-$1938 & 0031$-$196 & 144 & 2.34,0.93,$-$1.1 & 26 & 0.022 \\ [2pt] 1ES~0033+595 & 0033+595 & 132 & 1.61,0.84,0.7 & 43 & 0.024 \\ [2pt] RGB~J0136+391 & 0133+388 & 150 & 1.93,0.93,13.3 & 35 & 0.020 \\ [2pt] RGB~J0152+017 & 0150+015 & 96 & 2.12,0.92,0.9 & 43 & 0.025 \\ [2pt] 1ES~0229+200 & 0229+200 & 96 & 1.93,0.94,$-$0.2 & 21 & 0.023 \\ [2pt] RBS~0413 & 0317+185 & 96 & 1.89,0.94,1.3 & 18 & 0.025 \\ [2pt] 1ES~0347$-$121 & 0347$-$121 & 96 & 2.25,0.89,$-$0.9 & 7 & 0.025 \\ [2pt] 1ES~0414+009 & 0414+009 & 104 & 2.04,0.87,$-$1.7 & 35 & 0.022 \\ [2pt] 1ES~0502+675 & 0502+675 & 104 & 1.34,1.01,0.5 & 19 & 0.023 \\ [2pt] PKS~0548$-$322 & 0548$-$322 & 104 & 2.19,0.84,1.0 & 20 & 0.062 \\ [2pt] RX~J0648.7+1516 & 0645+153 & 160 & 1.92,0.86,$-$3.0 & 36 & 0.020 \\ [2pt] 1ES~0647+250 & 0647+251 & 160 & 1.88,0.88,$-$4.9 & 43 & 0.018 \\ [2pt] RGB~J0710+591 & 0706+592 & 104 & 1.42,1.03,14.5 & 28 & 0.023 \\ [2pt] 1RXS~J101015.9$-$311909 & 1008$-$310 & 120 & 2.20,0.81,$-$2.7 & 29 & 0.040 \\ [2pt] MS~1221.8+2452 & 1221+248 & 132 & 1.83,0.84,$-$0.1 & 16 & 0.023 \\ [2pt] 1ES~1440+122 & 1440+122 & 117 & 1.98,0.87,$-$7.2 & 18 & 0.025 \\ [2pt] H~1722+119 & 1722+119 & 117 & 2.04,0.98,$-$11.8 & 66 & 0.030 \\ [2pt] 1ES~1741+196 & 1741+196 & 117 & 1.95,0.97,$-$12.6 & 98 & 0.030 \\ [2pt] B3~2247+381 & 2247+381 & 117 & 2.11,0.81,2.5 & 42 & 0.029 \\ \tableline \\ [-10pt] \end{tabular}} \end{center} {\bf Notes.}\\ $a$: Numbers are for the naturally weighted beam, and are the FWHMs of the major and minor axes in mas, and the position angle of the major axis in degrees. Position angle is measured from north through east.\\ $b$: Rms noise in the total intensity image.\\ \end{table*} About half of these sources have been previously imaged with the VLBA by other investigators, although all of those images were obtained prior to the VLBA sensitivity upgrade. Rector et al.\ (2003) show 5~GHz VLBA images of the five sources 0033+595, 0229+200, 0414+009, 0647+251, and 1741+196. Those images have about twice the beam size and about three times the noise level of the images in Figure~1, but they all agree in showing the same general extended jet structure. Giroletti et al.\ (2004b) show 5~GHz VLBA images of the five sources 0229+200, 0347$-$121, 0548$-$322, 0706+592, and 1440+122; however, those images all have about six times the noise level of the images in Figure~1, and they only detect parsec-scale jet structure in 0706+592. The source 1722+119 has a single image in the MOJAVE database, but it shows only the VLBI core. Collectively, these prior imaging results for those ten sources demonstrate the importance of the VLBA sensitivity upgrade in imaging the parsec-scale structure of the TeV HBLs. For the remaining ten sources in Figure~1, these are the first published VLBI images known to the authors. The general parsec-scale morphology of the sources in Figure~1 is familiar from VLBI studies of brighter TeV blazars; for example, Mrk~421 and Mrk~501 (Piner et al.\ 1999; Edwards \& Piner 2002; Giroletti et al.\ 2006, 2008). Most of the sources show a collimated jet a few milliarcseconds long that transitions to a lower surface brightness, more diffuse jet with a broader opening angle at a few mas from the core. The structure at tens of milliarcseconds from the core at 8~GHz then appears patchy and filamentary. As an example, the source 0706+592 nicely displays this morphology in Figure~1. Despite this general pattern, there are a couple of sources with unusual morphologies. The sources 0033+595 and 0647+251 both show structure on opposite sides of the presumed core. Either the brightest most compact component is not the core, the jet crosses back over the line of sight (as seen in the TeV blazar 1ES~1959+650 by Piner et al.\ 2008), or the emission is truly two-sided. Forthcoming imaging at multiple frequencies should identify the core for these unusual cases. At least two sources (0502+675 and 1722+119) also display limb-brightened jets in their inner jet region, which is discussed further in $\S$~\ref{transverse}. \subsection{Model Fits} \label{mfits} After imaging and final calibration of the visibilities, we fit circular Gaussian models to the calibrated visibilities for each source using the {\em modelfit} task in DIFMAP. Circular Gaussians are more stable than elliptical Gaussians during fitting, and they provided adequate fits to the visibilities for all sources, as noted by the reduced chi-squared of the fit and visual inspection of the residual map and visibilities. Model fitting directly to the visibilities allows sub-beam resolution to be obtained, and components can be clearly identified in the model fitting even when they appear blended with the core component or with each other in the CLEAN images. In a number of cases patchy low surface brightness emission beyond the collimated jet region could not be well-fit by a circular Gaussian, so the model fits do not necessarily represent the most distant emission seen on the CLEAN images. Note also that, because of incomplete sampling in the $(u,v)$-plane, VLBI model fits are not unique, and represent only one mathematically possible deconvolution of the source structure. The circular Gaussian models fit to all 20 sources are given in Table~\ref{mfittab}. Note that flux values for closely spaced components may be inaccurate, since it is difficult for the fitting algorithm to uniquely distribute the flux during model fitting. The model component naming follows the scheme used in our previous papers (e.g., Piner et al.\ 2010); jet components are numbered C1, C2, etc., from the outermost component inward. Observer-frame brightness temperatures are also given in Table~\ref{mfittab} for all partially-resolved core components (those whose best-fit size is not zero). These VLBI core brightness temperatures and associated errors are discussed in detail in the following subsection. \begin{table*} [!h] \begin{center} {\normalsize \caption{Circular Gaussian Models} \label{mfittab} \begin{tabular}{l c c r r r c c c} \tableline \tableline \\ [-5pt] \multicolumn{1}{c}{Source} & B1950 & Component & \multicolumn{1}{c}{$S$} & \multicolumn{1}{c}{$r$} & \multicolumn{1}{c}{PA} & $a$ & $\chi_{R}^{2}$ & $T_{B}$ \\ \multicolumn{1}{c}{(1)} & Name & (3) & \multicolumn{1}{c}{(mJy)} & \multicolumn{1}{c}{(mas)} & \multicolumn{1}{c}{(deg)} & (mas) & (8) & ($10^{10}$ K) \\ & (2) & & \multicolumn{1}{c}{(4)} & \multicolumn{1}{c}{(5)} & \multicolumn{1}{c}{(6)} & (7) & & (9) \\ \tableline \\ [-5pt] SHBL~J001355.9$-$185406 & 0011$-$191 & Core & 9.3 & ... & ... & 0.23 & 0.66 & 0.3 \\ [2pt] & & C1 & 5.0 & 0.99 & $-$42.4 & 0.98 & & \\ [2pt] KUV~00311$-$1938 & 0031$-$196 & Core & 25.8 & ... & ... & 0.10 & 0.74 & 4.3 \\ [2pt] & & C2 & 2.7 & 2.55 & 29.9 & 1.39 & & \\ [2pt] & & C1 & 2.6 & 4.56 & 22.7 & 3.54 & & \\ [2pt] 1ES~0033+595 & 0033+595 & Core & 43.7 & ... & ... & 0.23 & 0.72 & 1.5 \\ [2pt] & & C2 & 9.7 & 1.18 & $-$24.1 & 1.16 & & \\ [2pt] & & C1 & 5.1 & 5.52 & 69.1 & 4.30 & & \\ [2pt] RGB~J0136+391 & 0133+388 & Core & 28.7 & ... & ... & 0.00 & 0.75 & ... \\ [2pt] & & C3 & 6.8 & 0.36 & $-$7.9 & 0.52 & & \\ [2pt] & & C2 & 2.3 & 0.97 & $-$59.6 & 1.92 & & \\ [2pt] & & C1 & 2.5 & 7.99 & $-$92.4 & 9.45 & & \\ [2pt] RGB~J0152+017 & 0150+015 & Core & 42.6 & ... & ... & 0.16 & 0.80 & 2.9 \\ [2pt] & & C2 & 4.8 & 1.02 & $-$135.6 & 0.68 & & \\ [2pt] & & C1 & 2.9 & 3.27 & $-$125.4 & 1.61 & & \\ [2pt] 1ES~0229+200 & 0229+200 & Core & 19.9 & ... & ... & 0.10 & 0.74 & 3.6 \\ [2pt] & & C4 & 2.2 & 0.94 & 163.6 & 0.50 & & \\ [2pt] & & C3 & 2.3 & 3.06 & 154.6 & 1.04 & & \\ [2pt] & & C2 & 1.4 & 6.71 & 160.8 & 2.80 & & \\ [2pt] & & C1 & 2.4 & 15.76 & 171.6 & 8.79 & & \\ [2pt] RBS~0413 & 0317+185 & Core & 16.7 & ... & ... & 0.08 & 0.74 & 4.5 \\ [2pt] & & C3 & 2.6 & 0.85 & $-$13.9 & 0.38 & & \\ [2pt] & & C2 & 1.1 & 2.12 & $-$14.8 & 0.70 & & \\ [2pt] & & C1 & 1.2 & 5.03 & $-$11.2 & 1.53 & & \\ [2pt] 1ES~0347$-$121 & 0347$-$121 & Core & 7.4 & ... & ... & 0.14 & 0.89 & 0.6 \\ [2pt] & & C1 & 1.6 & 1.74 & $-$144.2 & 1.38 & & \\ [2pt] 1ES~0414+009 & 0414+009 & Core & 35.7 & ... & ... & 0.28 & 0.78 & 0.8 \\ [2pt] & & C1 & 11.1 & 1.41 & 85.7 & 2.59 & & \\ [2pt] 1ES~0502+675 & 0502+675 & Core & 17.2 & ... & ... & 0.26 & 0.75 & 0.4 \\ [2pt] & & C2 & 2.9 & 0.37 & $-$139.0 & 0.41 & & \\ [2pt] & & C1 & 3.2 & 4.58 & $-$74.5 & 6.90 & & \\ [2pt] PKS~0548$-$322 & 0548$-$322 & Core & 20.4 & ... & ... & 0.32 & 0.69 & 0.3 \\ [2pt] & & C1 & 6.2 & 1.27 & 30.9 & 0.61 & & \\ [2pt] RX~J0648.7+1516 & 0645+153 & Core & 33.7 & ... & ... & 0.00 & 0.76 & ... \\ [2pt] & & C5 & 3.9 & 0.81 & $-$1.9 & 0.32 & & \\ [2pt] & & C4 & 2.1 & 2.43 & $-$1.6 & 0.87 & & \\ [2pt] & & C3 & 1.1 & 4.67 & $-$10.9 & 2.08 & & \\ [2pt] & & C2 & 2.6 & 10.77 & $-$1.3 & 5.09 & & \\ [2pt] & & C1 & 3.3 & 21.51 & 5.6 & 7.18 & & \\ [2pt] 1ES~0647+250 & 0647+251 & Core & 41.6 & ... & ... & 0.15 & 0.76 & 3.2 \\ [2pt] & & C2 & 9.1 & 0.91 & 157.9 & 1.78 & & \\ [2pt] & & C1 & 1.8 & 3.28 & $-$82.6 & 3.29 & & \\ [2pt] RGB~J0710+591 & 0706+592 & Core & 27.0 & ... & ... & 0.17 & 0.75 & 1.5 \\ [2pt] & & C3 & 4.8 & 0.80 & $-$145.3 & 0.63 & & \\ [2pt] & & C2 & 3.8 & 3.10 & $-$156.0 & 1.83 & & \\ [2pt] & & C1 & 4.9 & 14.50 & $-$156.7 & 10.57 & & \\ [2pt] \end{tabular}} \end{center} \end{table*} \begin{table*}[!t] \begin{center} {\normalsize Table 4 (cont.): Circular Gaussian Models \\ \begin{tabular}{l c c r r r c c c} \tableline \tableline \\ [-5pt] \multicolumn{1}{c}{Source} & B1950 & Component & \multicolumn{1}{c}{$S$} & \multicolumn{1}{c}{$r$} & \multicolumn{1}{c}{PA} & $a$ & $\chi_{R}^{2}$ & $T_{B}$ \\ \multicolumn{1}{c}{(1)} & Name & (3) & \multicolumn{1}{c}{(mJy)} & \multicolumn{1}{c}{(mas)} & \multicolumn{1}{c}{(deg)} & (mas) & (8) & ($10^{10}$ K) \\ & (2) & & \multicolumn{1}{c}{(4)} & \multicolumn{1}{c}{(5)} & \multicolumn{1}{c}{(6)} & (7) & & (9) \\ \tableline \\ [-5pt] 1RXS~J101015.9$-$311909 & 1008$-$310 & Core & 30.1 & ... & ... & 0.29 & 0.65 & 0.6 \\ [2pt] & & C1 & 5.1 & 0.91 & 32.7 & 2.01 & & \\ [2pt] MS~1221.8+2452 & 1221+248 & Core & 16.0 & ... & ... & 0.22 & 0.68 & 0.6 \\ [2pt] & & C2 & 2.6 & 0.77 & $-$129.5 & 0.65 & & \\ [2pt] & & C1 & 1.6 & 2.15 & $-$139.0 & 1.12 & & \\ [2pt] 1ES~1440+122 & 1440+122 & Core & 16.7 & ... & ... & 0.00 & 0.67 & ... \\ [2pt] & & C3 & 2.5 & 0.48 & $-$49.6 & 0.42 & & \\ [2pt] & & C2 & 1.5 & 1.68 & $-$71.8 & 1.47 & & \\ [2pt] & & C1 & 1.3 & 12.41 & $-$81.6 & 6.59 & & \\ [2pt] H~1722+119 & 1722+119 & Core & 62.5 & ... & ... & 0.00 & 0.68 & ... \\ [2pt] & & C2 & 3.7 & 0.58 & 152.4 & 0.00 & & \\ [2pt] & & C1 & 4.0 & 3.50 & 163.3 & 4.29 & & \\ [2pt] 1ES~1741+196 & 1741+196 & Core & 94.8 & ... & ... & 0.24 & 0.78 & 2.9 \\ [2pt] & & C3 & 22.0 & 0.75 & 70.0 & 0.67 & & \\ [2pt] & & C2 & 17.5 & 2.08 & 71.7 & 1.50 & & \\ [2pt] & & C1 & 7.0 & 6.00 & 81.1 & 2.43 & & \\ [2pt] B3~2247+381 & 2247+381 & Core & 39.8 & ... & ... & 0.16 & 0.75 & 2.7 \\ [2pt] & & C3 & 10.3 & 0.69 & $-$91.5 & 1.53 & & \\ [2pt] & & C2 & 3.4 & 4.63 & $-$74.6 & 1.70 & & \\ [2pt] & & C1 & 1.7 & 9.61 & $-$59.7 & 3.65 & & \\ \tableline \\ [-10pt] \end{tabular}} \end{center} {\bf Notes.} Column~4: flux density in millijanskys. Columns~5 and 6: $r$ and PA are the polar coordinates of the center of the component relative to the presumed core. Position angle is measured from north through east. Column~7: FWHM of the Gaussian component. Column~8: the reduced chi-squared of the model fit. Column~9: the maximum observer-frame brightness temperature of the Gaussian core component is given by $T_{B}=1.22\times10^{12}S/(a^{2}\nu^{2})$~K, where $S$ is the flux density in janskys, $a$ is the FWHM in mas, and $\nu$ is the observation frequency in GHz. Brightness temperature is given for core components whose best-fit size is not zero. \end{table*} \subsection{Core Brightness Temperatures} \label{tb} The observed brightness temperatures of VLBI cores can be used to constrain both Doppler beaming factors and the physical processes occurring in a source. The maximum observer-frame brightness temperature of a circular Gaussian is \begin{equation} \label{tbeq} T_{B}=1.22\times10^{12}\;\frac{S}{a^{2}\nu^{2}}\;\rm{K}, \end{equation} where $S$ is the flux density of the Gaussian in janskys, $a$ is the FWHM of the Gaussian in mas, and $\nu$ is the observing frequency in GHz. Note that we use observer-frame brightness temperatures, i.e., without applying the $(1+z)$ factor to convert to source-frame brightness temperatures, because the redshift of a number of these sources is uncertain (see Table~\ref{alltab}). The median redshift of the sources with known redshift is about 0.2, so source-frame brightness temperatures are only about 20\% higher. Several mechanisms can act to limit the intrinsic rest-frame brightness temperature of a synchrotron source; e.g., rapid energy loss by inverse Compton emission that limits the brightness temperature to $\sim5\times10^{11} - 1\times10^{12}$~K (Kellermann \& Pauliny-Toth 1969), or equipartition of energy between particles and magnetic fields that limits the brightness temperature to $\sim5\times10^{10} - 1\times10^{11}$~K (Readhead 1994). The observed brightness temperature of a source is increased relative to its intrinsic brightness temperature by a factor of the Doppler factor $\delta$ \footnote{The Doppler factor $\delta=1/(\gamma(1-\beta\cos\theta))$, where $\theta$ is the viewing angle, $\beta =v/c$, and $\gamma=(1-\beta^2)^{-1/2}$ is the bulk Lorentz factor.}. Thus, if upper limits to the intrinsic brightness temperature are known, then the observed brightness temperature can be used to compute a lower limit to the Doppler factor. Similarly, if the Doppler factor can be estimated by independent means, then intrinsic brightness temperatures can be computed from observed values. This has been done by, e.g., L\"{a}hteenm\"{a}ki et al.\ (1999), Homan et al.\ (2006), and Hovatta et al.\ (2013), who all used either apparent superluminal speeds or total flux density variability to compute intrinsic brightness temperatures of AGN samples. All of these studies concluded that typical intrinsic brightness temperatures were in the range of few times $10^{10}$ to $10^{11}$~K, and therefore likely to be limited by equipartition of energy. Care must be taken in the analysis of these VLBI core brightness temperatures, because many of the cores are only partially resolved, and even though a `best-fit' size may be returned by the model fitting routine, the fit is often nearly as good if the Gaussian component is simply replaced by a delta function. In such cases, only an upper limit to the size, or a lower limit to the brightness temperature, can actually be measured. The flux density of the core, and the baseline lengths and sensitivity of the VLBI array, determine the maximum measurable brightness temperature (e.g., Lovell et al.\ 2000; Wehrle et al.\ 2001; Kovalev et al.\ 2005; Lobanov 2005), which is of order $10^{11}$~K for these observations. Because some core brightness temperatures in Table~\ref{mfittab} are within factors of a few of this value, we conducted a full error analysis of the core brightness temperatures; this error analysis is described below and tabulated in Table~\ref{tbtab}. \begin{table*}[!t] \begin{center} {\normalsize \caption{Brightness Temperature Error Analysis} \label{tbtab} \begin{tabular}{c r c c r c c r c c} \tableline \tableline \\ [-5pt] B1950 & \multicolumn{1}{c}{$S$} & $a$ & $T_{B}$ & \multicolumn{1}{c}{$S_{\rm{max}}$} & $a_{\rm{min}}$ & $T_{B,\rm{max}}$ & \multicolumn{1}{c}{$S_{\rm{min}}$} & $a_{\rm{max}}$ & $T_{B,\rm{min}}$ \\ Name & \multicolumn{1}{c}{(mJy)} & (mas) & ($10^{10}$ K) & \multicolumn{1}{c}{(mJy)} & (mas) & ($10^{10}$ K) & \multicolumn{1}{c}{(mJy)} & (mas) & ($10^{10}$ K) \\ (1) & \multicolumn{1}{c}{(2)} & (3) & (4) & \multicolumn{1}{c}{(5)} & (6) & (7) & \multicolumn{1}{c}{(8)} & (9) & (10) \\ \tableline \\ [-5pt] 0011$-$191 & 9.3 & 0.23 & 0.3 & 13.4 & 0.00 & $\infty$ & 7.1 & 0.36 & 0.1 \\ [2pt] 0031$-$196 & 25.8 & 0.10 & 4.3 & 29.3 & 0.00 & $\infty$ & 22.3 & 0.19 & 1.1 \\ [2pt] 0033+595 & 43.7 & 0.23 & 1.5 & 48.2 & 0.14 & 4.4 & 39.2 & 0.30 & 0.8 \\ [2pt] 0133+388 & 28.7 & 0.00 & $\infty$ & 33.3 & 0.00 & $\infty$ & 22.3 & 0.27 & 0.5 \\ [2pt] 0150+015 & 42.6 & 0.16 & 2.9 & 45.1 & 0.09 & 10.3 & 38.1 & 0.23 & 1.3 \\ [2pt] 0229+200 & 19.9 & 0.10 & 3.6 & 23.4 & 0.00 & $\infty$ & 16.4 & 0.23 & 0.5 \\ [2pt] 0317+185 & 16.7 & 0.08 & 4.5 & 19.2 & 0.00 & $\infty$ & 14.2 & 0.22 & 0.5 \\ [2pt] 0347$-$121 & 7.4 & 0.14 & 0.6 & 10.9 & 0.00 & $\infty$ & 4.9 & 0.28 & 0.1 \\ [2pt] 0414+009 & 35.7 & 0.28 & 0.8 & 40.2 & 0.17 & 2.4 & 32.2 & 0.33 & 0.5 \\ [2pt] 0502+675 & 17.2 & 0.26 & 0.4 & 22.7 & 0.11 & 3.1 & 13.7 & 0.37 & 0.2 \\ [2pt] 0548$-$322 & 20.4 & 0.32 & 0.3 & 26.9 & 0.12 & 3.0 & 15.9 & 0.36 & 0.2 \\ [2pt] 0645+153 & 33.7 & 0.00 & $\infty$ & 39.0 & 0.00 & $\infty$ & 30.0 & 0.21 & 1.2 \\ [2pt] 0647+251 & 41.6 & 0.15 & 3.2 & 45.1 & 0.10 & 8.2 & 38.1 & 0.23 & 1.2 \\ [2pt] 0706+592 & 27.0 & 0.17 & 1.5 & 31.5 & 0.00 & $\infty$ & 22.5 & 0.29 & 0.5 \\ [2pt] 1008$-$310 & 30.1 & 0.29 & 0.6 & 37.6 & 0.13 & 3.9 & 24.6 & 0.39 & 0.3 \\ [2pt] 1221+248 & 16.0 & 0.22 & 0.6 & 20.5 & 0.00 & $\infty$ & 12.5 & 0.34 & 0.2 \\ [2pt] 1440+122 & 16.7 & 0.00 & $\infty$ & 20.2 & 0.00 & $\infty$ & 13.2 & 0.23 & 0.4 \\ [2pt] 1722+119 & 62.5 & 0.00 & $\infty$ & 65.9 & 0.00 & $\infty$ & 58.9 & 0.13 & 6.0 \\ [2pt] 1741+196 & 94.8 & 0.24 & 2.9 & 98.0 & 0.21 & 3.8 & 89.2 & 0.28 & 1.9 \\ [2pt] 2247+381 & 39.8 & 0.16 & 2.7 & 44.3 & 0.00 & $\infty$ & 35.3 & 0.23 & 1.1 \\ \tableline \\ [-10pt] \end{tabular}} \end{center} {\bf Notes.} Columns 2, 3, and 4 are the best-fit core flux density and size, and the associated observer's frame brightness temperature. These values are also given in Table~\ref{mfittab}. Columns 5, 6, and 7 are the maximum allowed core flux density and the minimum allowed size, and the associated maximum brightness temperature computed from those quantities. Columns 8, 9, and 10 are the minimum allowed core flux density and the maximum allowed size, and the associated minimum brightness temperature computed from those quantities. \end{table*} We used the Difwrap program (Lovell 2000), as described by e.g., Piner et al.\ (2000) and Tingay et al.\ (2001), to determine upper and lower bounds to the measured brightness temperatures. We established minimum and maximum values for the flux density and size of a component by systematically varying that property, while allowing other parameters to re-converge, and then visually comparing the new fit to the measured visibilities. The upper bound to the brightness temperature was then computed using the maximum flux and the minimum size, while the lower bound to the brightness temperature was computed using the minimum flux and the maximum size. All of these values are tabulated in Table~\ref{tbtab}. Roughly half of the core components are consistent with a size of zero, meaning that the associated brightness temperature measurements have no upper bound and are only lower limits, but the half that do have upper bounds provide valuable constraints. If we exclude the single high-brightness temperature source 1722+119 ($T_{B}>6.0\times10^{10}$~K), then the largest lower-limit is $T_{B}>1.9\times10^{10}$~K for 1741+196. Similarly, the smallest upper-limit is $T_{B}<2.4\times10^{10}$~K for 0414+009. Thus, all except one of these twenty sources are consistent with the brightness temperature range $1.9\times10^{10}<T_{B}<2.4\times10^{10}$~K, and we therefore take a brightness temperature of $\sim2\times10^{10}$~K to be a typical observed brightness temperature of a TeV HBL. This is consistent with typical observed brightness temperatures of TeV HBLs measured in earlier works (e.g., Piner et al.\ 2010), but it is now established for a much larger number of sources. To compare with intrinsic brightness temperature limits that have been derived for homogeneous optically thick spheres, we can convert our Gaussian brightness temperatures to homogeneous sphere brightness temperatures by multiplying by the appropriate correction factor of 0.56 (e.g., Pearson 1995; Tingay et al.\ 2001). This yields a value of about $1\times10^{10}$~K as a typical brightness temperature of a TeV HBL. Comparing the typical observed brightness temperatures to the equipartition brightness temperatures calculated for these sources of about $6\times10^{10}$~K (Equation~4$a$ of Readhead 1994), we see that the observed brightness temperatures of these TeV HBLs are already at or below the equipartition limit with no need to invoke high Doppler factors to reduce the observed brightness temperatures. There is thus no evidence of relativistic beaming of the core emission based on the VLBI brightness temperatures. Even with no Doppler boosting, the observed brightness temperatures are already somewhat below the equipartition value, placing these sources in the magnetically-dominated regime (Readhead 1994; Homan et al.\ 2006). High values of the Doppler factor would reduce the intrinsic brightness temperature even more, placing the sources even farther from equipartition. We note one important caveat: that both observed brightness temperatures and intrinsic physical limits are traditionally calculated for a homogeneous sphere geometry, and that the actual geometry for the VLBI core region may be more complex, such as a partially-resolved limb-brightened structure (see $\S$~\ref{transverse}). Despite the lack of evidence for beaming from the VLBI brightness temperatures, the one-sided core-jet morphology displayed by the majority of these sources does imply at least mild Doppler boosting. However, because the sources studied in this paper are relatively faint, this constraint is modest. We have computed lower limits to the jet-to-counterjet brightness ratio for each source, based on the peak jet brightness from the model fits in Table~\ref{mfittab} restored with the associated beam from Table~\ref{imtab}, and using three times the rms noise from Table~\ref{imtab} as the minimum detectable counterjet brightness. The median lower limit to the jet-to-counterjet brightness ratio is 37:1, which implies $\delta>2$ for viewing angles of a few degrees. The highest lower limit is 210:1 for 1741+196, which implies $\gamma>2$ and $\delta>4$ for viewing angles of a few degrees. \subsection{Opening Angles} \label{opening} We have calculated the apparent opening angle $\phi_{app}$ of each of these twenty jets, using the model fits from Table~\ref{mfittab}, and the model-fitting approach to measuring apparent opening angles described by Pushkarev et al.\ (2009). These apparent opening angles are tabulated in Table~\ref{optab}. Note that the apparent opening angle is a function of both the intrinsic opening angle and the viewing angle through the relation $\phi_{app}\approx\phi_{int}/\sin\theta$, where $\theta$ is the viewing angle. Pushkarev et al.\ (2009) compared apparent opening angles of {\em Fermi}-detected and non-detected blazars, and found a tendency for the {\em Fermi}-detected blazars to have wider apparent opening angles than the non-detected ones. Because the calculated intrinsic opening angles of the two groups were similar, they suggested that the {\em Fermi}-detected jets were viewed more closely to the line of sight. Lister et al.\ (2011) used a larger sample of {\em Fermi}-detected blazars to compute a mean apparent opening angle of 24$\arcdeg$ for this sample, and they also found a positive correlation between apparent opening angle and the gamma-ray loudness of the source. \begin{table*} \begin{center} {\normalsize \caption{Apparent Opening Angles} \label{optab} \begin{tabular}{c c c c c c c c} \tableline \tableline \\ [-5pt] B1950 & $\phi_{app}$ & B1950 & $\phi_{app}$ & B1950 & $\phi_{app}$ & B1950 & $\phi_{app}$ \\ Name & (deg) & Name & (deg) & Name & (deg) & Name & (deg)\\ \tableline \\ [-5pt] 0011$-$191 & 26.2 & 0229+200 & 13.0 & 0548$-$322 & 13.4 & 1221+248 & 18.9 \\ [2pt] 0031$-$196 & 18.2 & 0317+185 & 10.2 & 0645+153 & 11.3 & 1440+122 & 20.7 \\ [2pt] 0033+595 & 23.8 & 0347$-$121 & 21.6 & 0647+251 & 35.5 & 1722+119 & 31.5 \\ [2pt] 0133+388 & 37.2 & 0414+009 & 42.6 & 0706+592 & 19.3 & 1741+196 & 18.4 \\ [2pt] 0150+015 & 16.2 & 0502+675 & 33.1 & 1008$-$310 & 47.8 & 2247+381 & 23.0 \\ \tableline \\ [-10pt] \end{tabular}} \end{center} \end{table*} The apparent opening angles in Table~\ref{optab} range from 10$\arcdeg$ to 48$\arcdeg$, with a mean of $24\pm2\arcdeg$, identical to the mean found by Lister et al.\ (2011) for a larger sample of {\em Fermi}-detected blazars. Because 16 out of 20 of the TeV sources imaged for this paper are also {\em Fermi} sources (see Table~\ref{alltab}), this identical mean is not surprising, but it does show that the apparent opening angle distribution for a TeV HBL-selected subset of {\em Fermi} sources is similar to the overall {\em Fermi}-detected distribution. There is thus no evidence from apparent opening angles that TeV HBLs have different distributions of either viewing angle or intrinsic opening angle compared to the larger sample of Fermi sources studied by Lister et al.\ (2011). We find no significant correlation between the apparent opening angles in Table~\ref{optab} and TeV gamma-ray loudness (see $\S$~\ref{loudness}) as was found by Lister et al.\ (2011); however, this is not conclusive considering our small sample size of 20 sources. \subsection{Morphology and Transverse Jet Structure} \label{transverse} \begin{figure*}[!t] \centering \includegraphics[scale=0.65]{f2a.ps} \includegraphics[scale=0.65]{f2b.ps} \caption{Transverse brightness profiles showing limb-brightening for 0502+675 at 2~mas from the core (left panel), and for 1722+119 at 6~mas from the core (right panel). The rms noise levels (or uncertainty on the curves) are 0.023~mJy~bm$^{-1}$ for 0502+675, and 0.030~mJy~bm$^{-1}$ for 1722+119 (see Table~\ref{imtab}).} \end{figure*} AGN jets may be expected to develop transverse (so-called `spine-sheath' or `spine-layer') velocity structures on theoretical grounds (e.g., Henri \& Pelletier 1991), and the existence of these structures could explain some important observed properties of the TeV HBLs. For example, Ghisellini et al.\ (2005) consider how a jet with a low Lorentz factor layer and a high Lorentz factor spine could produce the discrepant Lorentz factors that are observed for TeV HBLs in the radio and gamma-ray, while at the same time the interaction between these two regions could serve to decelerate the spine. Recently, Tavecchio et al.\ (2014) have calculated whether spine-layer structures in the TeV HBLs could also produce high-energy PeV neutrinos, such as those detected by IceCube (Aartsen et al.\ 2014). Such spine-sheath structures might produce limb-brightening in the VLBI images of these jets if, for example, the jet bends away from the line-of-sight such that the low Lorentz factor layer acquires a higher Doppler factor than the high Lorentz factor spine (e.g., Giroletti et al.\ 2004a), or if the layer simply has a higher synchrotron emissivity in the radio than the spine (e.g., Sahayanathan 2009; Ghisellini et al.\ 2005). Note though that the presence of transverse intensity structures in VLBI images can have causes other than two-component outflows; for example, Clausen-Brown et al.\ (2011) show that limb-brightening can be observed for a uniform cylindrical jet with a helical magnetic field and no transverse structure under certain viewing geometries. Limb-brightening has been observed a number of times in VLBI images of two of the brightest and closest TeV HBLs: Mrk~421 and Mrk~501. Giroletti et al.\ (2004a, 2008), Piner et al.\ (2009), and Croke et al.\ (2010) have all reported limb-brightening in VLBI images of Mrk~501, at a wide variety of distances from the core. Similar results have been obtained for Mrk~421, both at lower frequencies (Giroletti et al.\ 2006) and at 43~GHz (Piner et al.\ 2010; Blasi et al.\ 2013). Transverse polarization structures, both in EVPA and fractional polarization, have also been observed in both of these sources, as well as in the TeV HBL 1ES~1959+650 (Piner et al.\ 2010). We have produced transverse brightness profiles for all 20 of the sources imaged for this paper, at numerous points along their jets. Many of these sources display the following general pattern in their transverse structure: the jets are well collimated and unresolved in the transverse direction for the first few milliarcseconds, after which they transition to patchy low surface brightness emission that is resolved but has numerous intensity peaks in a transverse brightness profile. Of the twenty sources, we see only two examples of a classic limb-brightening profile, in the sources 0502+675 and 1722+119. Transverse brightness profiles showing the limb-brightening for these two sources are shown in Figure~2; the limb-brightened structure of these two sources can also be seen directly on the images shown in Figure~1. For both of these sources, the limb brightening remains visible over a radial range of roughly 3~mas. We note that the absence of such a clear signature of limb-brightening in the other sources does not mean that such transverse intensity structure is non-existent in these jets. For example, the observations of limb-brightening by Piner et al.\ (2009) and Piner et al.\ (2010) in the relatively nearby TeV HBLs Mrk~501 and Mrk~421 were obtained from high-resolution 43~GHz observations, and after subtraction of the core and super-resolution of the jet in the transverse direction. That same linear scale would correspond to an angular separation that is well within the jet region that is transversely unresolved in the lower-resolution 8~GHz images of the more distant TeV HBLs presented in this paper. It is possible that the inner jets of these sources would display such structures if they could be transversely resolved with high-frequency VLBI; unfortunately, observations at high-frequency are much less sensitive, and all but the brightest few TeV HBLs are too faint for this. Whether `spine-sheath' structures are a nearly universal structure for TeV HBL jets is therefore ambiguous from these observations. \section{Results for the Entire TeV HBL Sample} \label{resultsall} \subsection{VLBI and Gamma-Ray Data for the Sample} \label{dataall} In this section, we combine the new VLBI data obtained in this paper with gamma-ray and VLBI data on the other TeV HBLs. The VLBI and gamma-ray properties of the 44 TeV HBLs currently listed in the TeVCat catalog are tabulated in Table~\ref{alltab}. We have attempted to quote a redshift value for every source in Table~\ref{alltab}, but in a number of cases (9 out of 44) these values are either uncertain or they are lower limits. These cases are clearly indicated in the notes to Table~\ref{alltab}, and those values should be used with caution. \begin{sidewaystable*} \begin{center} \vspace*{3.5in} {\footnotesize \caption{VLBI and Gamma-Ray Properties of the TeV HBLs} \label{alltab} \begin{tabular}{l c c c c c c c c c c c c c c c} \tableline \tableline \\ [-4pt] \multicolumn{1}{c}{Source} & $z$ & $\nu$ & VLBI & VLBI & $T_{B}$ & Ref & TeV & TeV & Cutoff & Ind & Ref & {\em Fermi} & Ind & Log & $\overline{m}$ \\ \multicolumn{1}{c}{(1)} & (2) & (GHz) & Total & Core & ($10^{10}$K) & (7) & Flux & Flux & (TeV) & (11) & (12) & Flux & (14) & $G_{\rm TeV}$ & (16) \\ & & (3) & (mJy) & (mJy) & (6) & & (Crab) & ($10^{-12}$ & (10) & & & ($10^{-12}$ & & (15) \\ & & & (4) & (5) & & & (8) & ph cm$^{-2}$s$^{-1}$) & & & & erg cm$^{-2}$s$^{-1}$) & & \\ & & & & & & & & (9) & & & & (13) & \\ [2pt] \tableline \\ [-2pt] SHBL~J001355.9$-$185406 & 0.095 & 8.4 & 15 & 9 & 0.3 & 1 & 0.006 & 0.8 & 0.31 & 3.40 & 1 & ... & ... & 2.89 & ... \\ [2pt] KUV~00311$-$1938 & 0.506$^{a}$ & 8.4 & 31 & 26 & 4.3 & 1 & 0.010 & 1.2$^{a}$ & 0.33 & 3.70 & 2 & 40.0 & 1.76 & 3.17 & ... \\ [2pt] 1ES~0033+595 & 0.240$^{a}$ & 8.4 & 59 & 44 & 1.5 & 1 & 0.015 & 5.5$^{b}$ & 0.15 & 3.80 & 3,4 & 28.6 & 1.87 & 2.36 & ... \\ [2pt] RGB~J0136+391 & 0.400$^{a}$ & 8.4 & 40 & 29 & ur & 1 & ... & ...$^{c}$ & ... & ... & 4 & 61.9 & 1.69 & ... & ... \\ [2pt] RGB~J0152+017 & 0.080 & 8.4 & 51 & 43 & 2.9 & 1 & 0.020 & 2.7 & 0.30 & 2.95 & 5 & 9.6 & 1.79 & 2.88 & ... \\ [2pt] 1ES~0229+200 & 0.140 & 8.4 & 28 & 20 & 3.6 & 1 & 0.017 & 2.3 & 0.30 & 2.59 & 6 & ... & ... & 3.19 & ... \\ [2pt] PKS~0301$-$243 & 0.266 & 15.4 & 218 & 157 & 5.2 & 2 & 0.014 & 3.3 & 0.20 & 4.60 & 7 & 76.6 & 1.94 & 1.83 & ... \\ [2pt] IC~310 & 0.019 & 15.4 & 102 & 62 & 2.0 & 2 & 0.023 & 3.1 & 0.30 & 2.00 & 8 & 9.8 & 2.10 & 2.86 & ... \\ [2pt] RBS~0413 & 0.190 & 8.4 & 22 & 17 & 4.5 & 1 & 0.009 & 1.5 & 0.25 & 3.18 & 9 & 17.0 & 1.55 & 2.88 & ... \\ [2pt] 1ES~0347$-$121 & 0.188 & 8.4 & 9 & 7 & 0.6 & 1 & 0.022 & 3.9 & 0.25 & 3.10 & 10 & ... & ... & 3.67 & ... \\ [2pt] 1ES~0414+009 & 0.287 & 8.4 & 49 & 36 & 0.8 & 1 & 0.021 & 5.2 & 0.20 & 3.40 & 11 & 7.8 & 1.98 & 2.88 & ... \\ [2pt] PKS~0447$-$439 & 0.200$^{b}$ & ... & ... & ... & ... & ... & 0.027 & 4.7 & 0.25 & 3.89 & 12 & 135.6 & 1.86 & ... & ... \\ [2pt] 1ES~0502+675 & 0.314$^{c}$ & 8.4 & 23 & 17 & 0.4 & 1 & 0.060 & 8.1$^{b}$ & 0.30 & 3.92 & 13 & 42.3 & 1.49 & 3.85 & ... \\ [2pt] PKS~0548$-$322 & 0.069 & 8.4 & 27 & 20 & 0.3 & 1 & 0.015 & 2.7 & 0.25 & 2.86 & 14 & ... & ... & 3.01 & ... \\ [2pt] RX~J0648.7+1516 & 0.179 & 8.4 & 43 & 34 & ur & 1 & 0.033 & 8.1$^{a}$ & 0.20 & 4.40 & 15 & 20.4 & 1.74 & 2.85 & ... \\ [2pt] 1ES~0647+250 & 0.450 & 8.4 & 53 & 42 & 3.2 & 1 & 0.030 & 19.5$^{b}$ & 0.10 & ... & 16 & 27.2 & 1.59 & 2.91 & ... \\ [2pt] RGB~J0710+591 & 0.125 & 8.4 & 40 & 27 & 1.5 & 1 & 0.029 & 3.9 & 0.30 & 2.69 & 17 & 13.3 & 1.53 & 3.23 & 0.064 \\ [2pt] 1ES~0806+524 & 0.138 & 22.2 & 89 & 64 & 0.9 & 3 & 0.016 & 2.2 & 0.30 & 3.60 & 18 & 27.9 & 1.94 & 2.53 & 0.120 \\ [2pt] RBS~0723 & 0.198 & 8.4 & 6 & 6 & ... & 4 & 0.025 & 6.1$^{b}$ & 0.20 & ... & 19 & 9.3 & 1.48 & 3.83 & ... \\ [2pt] 1RXS~J101015.9$-$311909 & 0.143 & 8.4 & 35 & 30 & 0.6 & 1 & 0.010 & 2.4 & 0.20 & 3.08 & 20 & 9.8 & 2.24 & 2.64 & ... \\ [2pt] 1ES~1011+496 & 0.212 & 15.4 & 191 & 106 & 2.5 & 3 & 0.065 & 15.8 & 0.20 & 4.00 & 21 & 72.6 & 1.85 & 2.59 & ... \\ [2pt] 1ES~1101$-$232 & 0.186 & 8.4 & 28 & 23 & 0.6 & 5 & 0.019 & 4.5 & 0.20 & 2.94 & 22 & 6.1 & 1.80 & 3.10 & ... \\ [2pt] Markarian~421 & 0.031 & 8.6 & 421 & 285 & 25.6 & 6 & 0.645 & 156.7$^{d}$ & 0.20 & 2.20 & 23 & 375.7 & 1.77 & 3.71 & 0.083 \\ [2pt] Markarian~180 & 0.045 & 22.2 & 83 & 40 & 7.6 & 5 & 0.093 & 22.5 & 0.20 & 3.30 & 24 & 14.9 & 1.74 & 3.10 & 0.094 \\ [2pt] RX~J1136.5+6737 & 0.134 & 8.4 & 23 & 19 & ur & 7 & 0.015 & 3.6$^{b}$ & 0.20 & ... & 25 & 8.4 & 1.68 & 2.97 & ... \\ [2pt] 1ES~1215+303 & 0.130 & 15.4 & 295 & 224 & ur & 2 & 0.032 & 7.7 & 0.20 & 2.96 & 26 & 61.2 & 2.02 & 2.26 & 0.087 \\ [2pt] 1ES~1218+304 & 0.184 & 8.4 & 39 & 24 & 1.5 & 5 & 0.050 & 12.2 & 0.20 & 3.08 & 27 & 37.8 & 1.71 & 3.34 & ... \\ [2pt] MS~1221.8+2452 & 0.218 & 8.4 & 21 & 16 & 0.6 & 1 & 0.040 & 9.7$^{b}$ & 0.20 & ... & 28 & 7.0 & 2.03 & 3.50 & ... \\ [2pt] 1ES~1312$-$423 & 0.105 & ... & ... & ... & ... & ... & 0.007 & 1.1$^{a}$ & 0.28 & 2.85 & 29 & ... & ... & ... & ... \\ [2pt] \end{tabular}} \end{center} \end{sidewaystable*} \begin{sidewaystable*} \begin{center} \vspace*{3.5in} {\footnotesize Table 7 (cont.): VLBI and Gamma-Ray Properties of the TeV HBLs \\ \begin{tabular}{l c c c c c c c c c c c c c c c} \tableline \tableline \\ [-4pt] \multicolumn{1}{c}{Source} & $z$ & $\nu$ & VLBI & VLBI & $T_{B}$ & Ref & TeV & TeV & Cutoff & Ind & Ref & {\em Fermi} & Ind & Log & $\overline{m}$ \\ \multicolumn{1}{c}{(1)} & (2) & (GHz) & Total & Core & ($10^{10}$K) & (7) & Flux & Flux & (TeV) & (11) & (12) & Flux & (14) & $G_{\rm TeV}$ & (16) \\ & & (3) & (mJy) & (mJy) & (6) & & (Crab) & ($10^{-12}$ & (10) & & & ($10^{-12}$ & & (15) \\ & & & (4) & (5) & & & (8) & ph cm$^{-2}$s$^{-1}$) & & & & erg cm$^{-2}$s$^{-1}$) & & \\ & & & & & & & & (9) & & & & (13) & \\ [2pt] \tableline \\ [-2pt] PKS~1424+240 & 0.604$^{a}$ & 15.4 & 218 & 123 & 6.1 & 2 & 0.042 & 21.0 & 0.12 & 3.80 & 30 & 145.0 & 1.78 & 2.41 & 0.069 \\ [2pt] H~1426+428 & 0.129 & 8.4 & 22 & 19 & 1.1 & 8 & 0.136 & 20.4 & 0.28 & 3.55 & 31 & 16.8 & 1.32 & 4.02 & ... \\ [2pt] 1ES~1440+122 & 0.163 & 8.4 & 22 & 17 & ur & 1 & 0.010 & 2.4$^{b}$ & 0.20 & 3.40 & 13 & 8.9 & 1.41 & 2.79 & ... \\ [2pt] PG~1553+113 & 0.500$^{d}$ & 22.2 & 134 & 95 & 1.2 & 5 & 0.080 & 29.3$^{b}$ & 0.15 & 4.27 & 32 & 197.4 & 1.67 & 2.87 & 0.082 \\ [2pt] Markarian~501 & 0.034 & 8.3 & 902 & 476 & 51.9 & 9 & 0.229 & 31.1$^{d}$ & 0.30 & 2.72 & 33 & 114.4 & 1.74 & 2.70 & 0.037 \\ [2pt] H~1722+119 & 0.170$^{a}$ & 8.4 & 70 & 63 & ur & 1 & 0.020 & 8.1$^{b}$ & 0.14 & ... & 34 & 42.3 & 1.93 & 2.52 & 0.114 \\ [2pt] 1ES~1727+502 & 0.055 & 15.4 & 101 & 68 & 5.8 & 2 & 0.021 & 7.7$^{b}$ & 0.15 & 3.20 & 35,4 & 9.7 & 1.83 & 2.31 & ... \\ [2pt] 1ES~1741+196 & 0.083 & 8.4 & 145 & 95 & 2.9 & 1 & 0.008 & 1.4$^{b}$ & 0.25 & ... & 36 & 9.4 & 1.62 & 1.93 & ... \\ [2pt] HESS~J1943+213 & 0.140$^{e}$ & 1.6 & 31 & 31 & 0.006 & 10 & 0.018 & 1.3 & 0.47 & 3.10 & 37 & ... & ... & 3.21 & ... \\ [2pt] 1ES~1959+650 & 0.047 & 15.4 & 150 & 91 & 2.3 & 11 & 0.146 & 19.8 & 0.30 & 2.72 & 38 & 66.9 & 1.94 & 3.30 & 0.115 \\ [2pt] PKS~2005$-$489 & 0.071 & 8.6 & 461 & 454 & 1.6 & 7 & 0.029 & 2.6 & 0.40 & 3.20 & 39 & 48.1 & 1.78 & 2.13 & ... \\ [2pt] PKS~2155$-$304 & 0.116 & 15.4 & 181 & 139 & 2.2 & 11 & 0.178 & 43.2 & 0.20 & 3.53 & 40 & 282.8 & 1.84 & 3.05 & ... \\ [2pt] B3~2247+381 & 0.119 & 8.4 & 55 & 40 & 2.7 & 1 & 0.021 & 5.0 & 0.20 & 3.20 & 41 & 13.2 & 1.84 & 2.72 & ... \\ [2pt] 1ES~2344+514 & 0.044 & 15.4 & 118 & 87 & 5.4 & 11 & 0.078 & 10.6 & 0.30 & 2.78 & 42 & 20.6 & 1.72 & 3.11 & ... \\ [2pt] H~2356$-$309 & 0.165 & 8.4 & 24 & 17 & 4.4 & 5 & 0.016 & 3.1 & 0.24 & 3.06 & 43 & 7.1 & 1.89 & 3.10 & ... \\ [2pt] \tableline \end{tabular}} \end{center} {\footnotesize {\bf Notes.} Column~1: `Canonical Name' from TeVCat, Column~2: redshift, Column~3: VLBI observing frequency, Column~4: total VLBI flux density, Column~5: VLBI core flux density, Column~6: core observer-frame Gaussian brightness temperature (ur=unresolved), Column~7: reference for VLBI data, Column~8: integrated TeV photon flux above the cutoff energy in multiples of the Crab flux, Column~9: integrated TeV photon flux above the cutoff energy, Column~10: cutoff energy for TeV flux, Column~11: TeV photon spectral index, Column~12: reference for TeV data, Columns~13 and 14: 2FGL {\em Fermi} energy flux and photon spectral index. Column~15: log TeV loudness (see $\S$~\ref{loudness}). Column~16: intrinsic modulation index from Richards et al.\ (2014) (see $\S$~\ref{modulation}).}\\ {\footnotesize {\bf Notes for Column 2.} $a$: Value is a lower limit. $b$: Uncertain. Value used is from Prandini et al.\ 2012. $c$: Uncertain. Value used is from NED, but is controversial. $d$: range 0.43 to 0.58. We use the mean. $e$: Value is a lower limit, but nature of source is controversial. See discussion later in text.}\\ {\footnotesize {\bf Notes for Column 9.} $a$: Computed from a differential flux from the reference. $b$: Computed from a flux in Crabs from the reference. $c$: Positive detection reported, no other information. $d$: Mean value computed from multiple fluxes in the reference.}\\ {\footnotesize {\bf References for Column 7.} (1) This paper; (2) MOJAVE program; (3) Piner \& Edwards 2013; (4) Bourda et al.\ 2010; (5) Tiet et al.\ 2012; (6) Piner et al.\ 2012; (7) {\url http://astrogeo.org/}; (8) Piner et al.\ 2008; (9) Piner et al.\ 2007; (10) Gab{\'a}nyi et al.\ 2013; (11) Piner \& Edwards 2004}\\ {\footnotesize {\bf References for Column 12.} (1) Abramowski et al.\ 2013a; (2) Becherini et al.\ 2012; (3) Mariotti 2011; (4) Mazin 2012; (5) Aharonian et al.\ 2008; (6) Aliu et al.\ 2014; (7) Abramowski et al.\ 2013b; (8) Aleksi{\'c} et al.\ 2010; (9) Aliu et al.\ 2012a; (10) Aharonian et al.\ 2007a; (11) Aliu et al.\ 2012b; (12) Abramowski et al.\ 2013d; (13) Benbow 2011; (14) Aharonian et al.\ 2010; (15) Aliu et al.\ 2011; (16) De Lotto 2012; (17) Acciari et al.\ 2010; (18) Acciari et al.\ 2009a; (19) Mirzoyan 2014a; (20) Abramowski et al.\ 2012; (21) Albert et al.\ 2007a; (22) Aharonian et al.\ 2007b; (23) Albert et al.\ 2007b; (24) Albert et al.\ 2006a; (25) Mirzoyan 2014b; (26) Aleksi{\'c} et al.\ 2012a; (27) Acciari et al.\ 2009b; (28) Cortina 2013a; (29) Abramowski et al.\ 2013c (30) Archambault et al.\ 2014; (31) Horan et al.\ 2002; (32) Aleksi{\'c} et al.\ 2012b; (33) Acciari et al.\ 2011b; (34) Cortina 2013b; (35) Aleksi{\'c} et al.\ 2014b; (36) Berger 2011; (37) Abramowski et al.\ 2011; (38) Albert et al.\ 2006b; (39) Acero et al.\ 2010; (40) Abramowski et al.\ 2010a; (41) Aleksi{\'c} et al.\ 2012c; (42)Acciari et al.\ 2011a; (43) Abramowski et al.\ 2010b} \end{sidewaystable*} About half of the VLBI data in Table~\ref{alltab} (20 sources) comes from this paper; VLBI data for most of the other HBLs comes from either our prior publications (13 sources), or from the MOJAVE program (5 sources). Four sources have VLBI data taken from elsewhere in the literature. Only two of the 44 sources (both below $-40\arcdeg$ declination) have no VLBI data in the literature. All of the VLBI data taken from elsewhere, such as that taken from the MOJAVE survey, has been independently model-fit by us if the visibility data files were available online; if not then published values have been used. For sources with multiple epochs of VLBI data, we have model-fit all epochs, and then used the epoch having the median core brightness temperature. VLBI data at an observing frequency of 8~GHz (the observing frequency used for this paper) was preferred if it was available; if not, then data at either 15 or 22 ~GHz (or in a single case, 1.6~GHz) has been used. References for all VLBI data used are given in the notes to Table~\ref{alltab}. The {\em Fermi} gamma-ray fluxes and spectral indices in Table~\ref{alltab} are taken from the 2FGL catalog (Nolan et al.\ 2012). Only six of the 44 sources have not been detected by {\em Fermi}, as of the 2FGL catalog. For the TeV gamma-ray data, an integrated photon flux, cutoff energy for that flux, and spectral index (uncorrected for Extra-galactic Background Light [EBL] absorption) was taken from the literature, if available. That integrated photon flux was then independently converted to a multiple of the Crab nebula flux using the Crab spectrum from Aharonian et al.\ (2006); this may cause slight differences from Crab fluxes quoted in the original papers referenced in Table~\ref{alltab}. In some cases, only a flux that was already expressed in multiples of the Crab flux was given in the literature. In those cases, the integrated photon flux above the cutoff energy was calculated from that using the Crab spectrum from Aharonian et al.\ (2006). For many of the sources, the numbers given in Table~\ref{alltab} match the flux in multiples of the Crab flux, cutoff energy, and spectral index quoted for that source in TeVCat; but in a number of cases they differ, due either to different literature sources used, or differing Crab nebula standards. Many of the newly discovered sources have only a single flux value in the literature, but for frequently observed sources, the variability of the TeV HBLs makes selection of a single flux value problematic. Because sources are observed over different time and energy ranges with different instruments, calculation of a formal mean would be difficult. Nevertheless, we have tried to select a `typical' flux value for variable sources, excluding extreme high or low states. This exclusion of extreme high or low states may also cause numbers in Table~\ref{alltab} to differ from those in TeVCat. In any event, the TeV data will almost certainly not be contemporaneous with the VLBI measurements. References for all TeV data used are given in the notes to Table~\ref{alltab}. \subsection{VLBI Flux Densities and Brightness Temperatures} \label{sandtb} A histogram of the VLBI core flux densities of the TeV HBLs from Table~\ref{alltab}, which is indicative of the most compact emission from these sources, is shown in Figure~3. New sources with VLBI data from this paper are shown in yellow, while sources with data taken from elsewhere are shown in blue. The range in core flux densities spans from a few mJy (e.g., 1ES~0347$-$121) to a few hundred mJy (e.g., Mrk~421 and Mrk~501), with a median of 38 mJy. Note that all of the new sources added in this paper have cores that are under 100 mJy. As the TeV gamma-ray telescopes have become more sensitive and begun to detect fainter objects, these sources have also tended to be fainter in the radio, a potential correlation that is explored in $\S$~\ref{correlations}. \begin{figure*}[!t] \centering \includegraphics[angle=90,scale=0.45]{f3.ps} \caption{Histogram of VLBI core flux densities of TeV HBLs from Table~\ref{alltab}. New sources with VLBI data from this paper are shown in yellow (20 sources). Sources with data taken from elsewhere are shown in blue (22 sources).} \end{figure*} \begin{figure*}[!ht] \centering \includegraphics[angle=90,scale=0.45]{f4.ps} \caption{Histogram of observer-frame Gaussian core brightness temperatures of TeV HBLs from Table~\ref{alltab}, for sources whose best-fit core size is not zero. New sources with VLBI data from this paper are shown in yellow (16 sources). Sources with data taken from elsewhere are shown in blue (19 sources). The outlier is HESS~J1943+213 (see text).} \end{figure*} A histogram of the VLBI core brightness temperatures from Table~\ref{alltab} is shown in Figure~4. New sources with VLBI data from this paper are shown in yellow, while sources with data taken from elsewhere are shown in blue. A brightness temperature value has been plotted in Figure~4 unless the best-fit value for the core size is zero (indicated by `ur' in Table~\ref{alltab}). However, as indicated by the brightness temperature error analysis done for the 20 sources observed for this paper in $\S$~\ref{tb}, some of these brightness temperature values are probably actually lower limits. The median brightness temperature in Figure~4 is $2\times10^{10}$~K, which is the same as the typical brightness temperature obtained from the brightness temperature error analysis in $\S$~\ref{tb}. See $\S$~\ref{tb} for discussion of the physical interpretation of such brightness temperatures in terms of intrinsic brightness temperature limits and relativistic beaming. Some outliers are notable in Figure~4. The only two TeV HBLs with brightness temperatures over $10^{11}$~K are the well-studied sources Mrk~421 and Mrk~501. The low brightness temperature outlier, with a measured brightness temperature of only $6\times10^{7}$~K, is the source HESS~J1943+213 which lies close to the Galactic plane. This source was observed with the European VLBI Network (EVN) at 1.6~GHz by Gab{\'a}nyi et al.\ (2013), who measured it to have a flux density of 31 mJy and an angular size of 16 mas, giving it a brightness temperature two orders of magnitude lower than all other TeV HBLs in Figure~4. The distribution in Figure~4 casts significant doubt on the HBL classification of this object (unless it is affected by an unusually large amount of interstellar scattering, see the discussion in Gab{\'a}nyi et al.\ 2013), and Gab{\'a}nyi et al.\ (2013) suggest instead a galactic origin for this source, in the form of a remote pulsar wind nebula (PWN). This interpretation may be strengthened by the lack of detection of any significant variability from this object from radio to TeV gamma-rays (Abramowski et al.\ 2011). \subsection{TeV Loudness} \label{loudness} In this section, we quantify the distribution of the ratio of TeV gamma-ray to radio luminosity present in the TeV HBL population. Lister et al.\ (2011) performed a similar analysis for {\em Fermi}-detected blazars by defining a quantity that they called the gamma-ray loudness, $G_{r}$. This quantity was defined by Lister et al.\ (2011) as the ratio of the gamma-ray luminosity between 0.1~GeV and 100~GeV, divided by the radio luminosity over a 15~GHz wide bandwidth, calculated from the VLBA flux density at 15~GHz (see equations 2 through 4 of Lister et al.\ 2011). We make straightforward modifications to Equations (2) though (4) of Lister et al.\ (2011) to adapt their gamma-ray loudness statistic to the TeV energy range and main VLBA observing frequency considered in this paper. We calculate the TeV loudness, $G_{\rm TeV}=L_{\rm TeV}/L_{\rm R}$, using the gamma-ray luminosity between 0.3 and 30~TeV, accounting for the different lower energy thresholds for the different measurements given in Table~\ref{alltab}. The modified versions of Equations (2) though (4) from Lister et al.\ (2011) are: \begin{equation} S_{\rm TeV}=\frac{(\Gamma-1)C_{1}E_{0}F_{0}}{(\Gamma-2)}\Bigg(\frac{E_{0}}{E_{1}}\Bigg)^{\Gamma-2}\Bigg[1-\Bigg(\frac{E_{1}}{E_{2}}\Bigg)^{\Gamma-2}\Bigg], \end{equation} where $F_{0}$ is the measured photon flux above the cutoff energy $E_{0}$, $\Gamma$ is the photon spectral index, $E_{1}=0.3$~TeV, $E_{2}=30$~TeV, and $C_{1}=1.602$~erg~TeV$^{-1}$, and $S_{\rm TeV}$ is in erg~cm$^{-2}$~s$^{-1}$, \begin{equation} L_{\rm TeV}=\frac{4\pi D_{L}^{2}S_{\rm TeV}}{(1+z)^{2-\Gamma}}~\rm{erg~s^{-1},} \end{equation} where $D_{L}$ is the luminosity distance in cm, and \begin{equation} L_{\rm R}=\frac{4\pi D_{L}^{2}\nu S_{\nu}}{(1+z)}~\rm{erg~s^{-1},} \end{equation} where $S_{\nu}$ is the total VLBA flux density in erg~cm$^{-2}$~s$^{-1}$~GHz$^{-1}$, and $\nu=8$~GHz. The quantities $F_{0}$, $E_{0}$, $\Gamma$, $S_{\nu}$, and $z$ are tabulated in Table~\ref{alltab}. If a photon spectral index was not measured for a source, then we used the median measured photon spectral index of $\Gamma=3.2$. We assume a flat radio spectral index ($\alpha=0$) for the radio $k$-correction and luminosity calculation. The TeV loudness is tabulated in Table~\ref{alltab}, and a histogram of this statistic is shown in Figure~5. As can be seen from Figure~5, the distribution spans about two orders of magnitude, from about 10$^{2}$ to 10$^{4}$, but the distribution is peaked around the median value of about 10$^{3}$. A similar range of about two orders of magnitude in gamma-ray loudness is spanned by the BL~Lac objects studied by Lister et al.\ (2011), although much of the range in gamma-ray loudness observed by those authors is due to a mix of HBLs, IBLs, and LBLs in the MOJAVE sample. The fact that a similar range is observed here among just the TeV HBLs is mostly due to the inclusion of the radio-faintest TeV HBLs at flux-density levels of a few millijanskys. For example, the single source with TeV loudness greater than 10$^{4}$ in Figure~5 is the extreme blazar H~1426+428, which is among the brighter TeV sources, but has a VLBA flux density of only about 20~mJy. Conversely, the two sources with TeV loudness less than 10$^{2}$ in Figure~5 are the relatively radio-bright HBLs PKS~0301$-$243 and 1ES~1741+196. \begin{figure*}[!t] \centering \includegraphics[angle=90,scale=0.45]{f5.ps} \caption{Histogram of the TeV loudness for the 41 TeV HBLs from Table~\ref{alltab} with both TeV and VLBI fluxes. The TeV loudness is defined in $\S$~\ref{loudness}.} \end{figure*} We might expect there to be a significant anti-correlation between TeV loudness and redshift, because of EBL absorption of TeV gamma-rays from distant sources. However, a correlation analysis of these two quantities does not yield a significant correlation, possibly because the vast majority of the TeV HBLs are clustered at low redshifts, and these low redshift sources already show a large intrinsic scatter in TeV loudness. Lister et al.\ (2011) found a significant anti-correlation between gamma-ray loudness and {\em Fermi} photon spectral index for the BL~Lac objects in their sample. We confirm this correlation for the TeV HBLs in this paper at a marginally significant level; for the 35 sources in Table~\ref{alltab} with both measured TeV loudness and photon index a partial Spearman rank correlation test (excluding effects of redshift) has a significance of 0.03. See Lister et al.\ (2011) for a discussion of the implications of such a correlation for emission models in BL~Lac objects. \subsection{Flux-Flux Correlations} \label{correlations} Establishing whether or not blazar fluxes in different wavebands (e.g., radio and gamma-ray) are intrinsically correlated, independent of any common-distance effects, is important to establishing to what degree the emission regions in the jet at these different wavebands are connected. The existence of correlations implies that the emission regions, even if they occur in different components of the jet with different beaming parameters, are related through some physical property of the source. Truly uncorrelated fluxes would instead imply that the emission regions probed by radio and gamma-ray observations are completely independent of each other. \begin{figure*}[!t] \centering \includegraphics[scale=0.65]{f6a.ps} \includegraphics[scale=0.65]{f6b.ps} \caption{Top panel: Plot of TeV flux in milliCrabs vs. total VLBA flux density in millijanskys (41 sources). Bottom panel: Plot of 2FGL {\em Fermi} flux vs. total VLBA flux density in millijanskys (37 sources). All flux values are from Table~\ref{alltab}.} \end{figure*} The top panel of Figure~6 shows the TeV flux in milliCrabs versus the total VLBI flux density in millijanskys for the 41 sources in Table~\ref{alltab} with measured TeV and VLBI fluxes. A partial correlation analysis (excluding effects of redshift) gives a Pearson partial correlation coefficient of 0.50, with a significance of $9\times10^{-4}$ (99.91\% chance of correlation). Repeating the analysis using the VLBI model-fit core flux density instead of the total flux density yields a similar but slightly lower Pearson partial correlation coefficient of 0.48, with a significance of $1.7\times10^{-3}$ (99.83\% chance of correlation). The correlation with core flux density is probably slightly less significant because the extra step of model fitting introduces some scatter into the core flux density values, particularly when there is a bright jet component close to the core. The high value for the significance of the correlation shown in the top panel of Figure~6 is partly due to the two sources Mrk~421 and Mrk~501, which are bright in both the radio and TeV gamma-rays. However, even if those two sources are excluded (and note that there is no particular physical reason for excluding them), the partial correlation remains significant, although at a lower level of 96.3\%. This is the first time, to our knowledge, that a correlation between the TeV flux and the VLBI flux has been established for the TeV HBL population. To further test the robustness of this new correlation, we employed Monte Carlo simulations using the method described by Pavlidou et al.\ (2012) to generate intrinsically uncorrelated permutations of the data for comparison purposes, using the monochromatic flux density at 0.3~TeV computed from Table~\ref{alltab} as the TeV flux sample. Because this method requires applying {\em k}-corrections to the permuted data, the six sources without a TeV photon spectral index in Table~\ref{alltab} were excluded. A comparison of $10^{7}$ randomly permuted datasets for the 35 remaining sources with the actual dataset yields a significance of correlation of 0.056 (94.4\% chance of correlation). This result is now only marginally significant; however, Pavlidou et al.\ (2012) state that their method is conservative for small samples such as this, and that existing intrinsic correlations may not be verified. We also note that the non-contemporaneous nature of the data may wash out a stronger correlation that might have existed in concurrently measured data. The bottom panel of Figure~6 shows the 2FGL {\em Fermi} flux versus the total VLBI flux density for the 37 sources in Table~\ref{alltab} with measured {\em Fermi} and VLBI fluxes. A similar plot is shown for the MOJAVE program sources in Figure~1 of Lister et al.\ (2011), and the bottom panel of Figure~6 basically continues the trend for the HBLs shown in that figure toward lower VLBI and {\em Fermi} flux values. Correlations between radio and {\em Fermi} fluxes for larger samples of {\em Fermi} blazars have been established by a number of authors (e.g., Kovalev et al.\ 2009; Ackermann et al.\ 2011; Linford et al.\ 2012). A correlation between {\em Fermi} and radio fluxes solely for the TeV HBL sub-population of {\em Fermi} sources was claimed by Xiong et al.\ (2013), although they did not address common-distance effects. We confirm such a correlation between the {\em Fermi} and VLBI fluxes of the TeV HBLs; a partial correlation analysis (excluding effects of redshift) gives a Pearson partial correlation coefficient of 0.76, with a formal significance of about $10^{-7}$. However, we note that because this sample is TeV-selected rather than {\em Fermi}-selected, such tests may overestimate the significance of correlations because they do not address upper limits for the TeV HBLs that are not in the 2FGL catalog, although this is a small number of objects (6 sources). \subsection{Radio Variability and Modulation Indices} \label{modulation} The variability of a radio source is an important property (potentially constraining both relativistic beaming and the relative locations of emission regions at different wavebands) that cannot be well-studied by sequences of a only a few VLBA images. The largest current effort to study radio variability of blazars is the Owens Valley Radio Observatory (OVRO) monitoring program\footnote{\url{http://www.astro.caltech.edu/ovroblazars}}, which presently monitors more than 1800 blazars about twice per week. The OVRO program's chosen parameter to characterize variability is the `intrinsic modulation index', $\overline{m}$, which is an estimate of the standard deviation of the source flux density divided by its mean (Richards et al.\ 2011, 2014). Although many of the TeV HBLs are in the OVRO program, because of their relative faintness they are clustered near the program's measurement limits in flux density and modulation index. For example, although 25 of the TeV HBLs in Table~\ref{alltab} are in Table~1 of Richards et al.\ (2014), 6 do not have a measured modulation index, and another 9 have a flux density and modulation index that are excluded from the analysis for being too close to the measurement limits. The 10 remaining high-confidence modulation indices are tabulated with the other source data in Table~\ref{alltab}. Because only 10 TeV HBLs pass the current data cuts in the OVRO analysis, we do not attempt correlation studies with the modulation index here, but leave it as an interesting possibility for future study if thresholds for the OVRO variability analysis are lowered. \section{Discussion and Conclusions} \label{discussion} We have investigated the parsec-scale jet structure of twenty relatively newly discovered TeV HBLs that had not been previously well-studied with VLBI. These newly discovered TeV HBLs extend down to only a few millijanskys in flux density, so they are not present in large VLBI monitoring programs. All sources were detected and imaged, and all showed parsec-scale jets that could be modeled with at least one Gaussian component ($\S$~\ref{mfits}). Most sources had a one-sided core-jet morphology, although we find two cases of apparently two-sided structure ($\S$~\ref{images}). Many sources show a common morphology of a collimated jet a few milliarcseconds long that transitions to a lower surface brightness, more diffuse jet with a broader opening angle at a few mas from the core. These results show that the entire TeV HBL sample, although relatively faint in the radio, is accessible to analysis with current VLBI instruments. As well as can be determined from only single-epoch images, the analyses presented here support previous conclusions that Lorentz factors in the parsec-scale cores and jets of TeV HBLs are only modestly relativistic. We determined allowed brightness temperature ranges for each core component ($\S$~\ref{tb}), and found that roughly half of the VLBI cores are resolved with brightness temperature upper limits of a few times $10^{10}$~K (Table~\ref{tbtab}). A Gaussian brightness temperature of $2\times10^{10}$~K was consistent with the data for all but one of the sources. Such brightness temperatures do not require any relativistic beaming to reduce them below likely intrinsic limits. The lack of detection of counter-jets does place at least a modest limit on the bulk Lorentz factor, although the strongest such constraint we could place was $\gamma\gtrsim2$. The distribution of apparent opening angles ($\S$~\ref{opening}) is indistinguishable from that of the general gamma-ray blazar population (Pushkarev et al.\ 2009; Lister et al.\ 2011), so there is no indication from their jet morphology that these sources are unusually close to the line-of-sight compared to other gamma-ray blazars. There is thus no evidence from these images that the slow apparent speeds of TeV HBLs are caused by a much closer alignment to the line-of-sight compared to the apparently faster sources. The `Doppler Crisis' for TeV HBLs suggests that their parsec-scale jets are structurally more complex than those of the more powerful blazars, and that they require at least two zones of significantly different Lorentz factor to successfully describe them. A consistent picture is emerging of this dichotomy in the jetted AGN population based on multiwavelength studies of large populations, theoretical modeling, and high-resolution imaging with VLBI. In this picture, jets formed in a low-efficiency accretion mode typical of HBLs (Ghisellini et al.\ 2005, 2009; Meyer et al.\ 2013b) favor interaction of the jet walls with the external medium, causing the formation of a slow layer. Radiative interaction between the spine and the layer may then decelerate the spine (Ghisellini et al.\ 2005), producing longitudinal as well as transverse velocity structure, such as postulated by Georganopoulos \& Kazanas (2003). Application of such two-zone models can also be successful in reducing the most extreme Doppler factors sometimes required by one-zone models. For example, for the TeV blazar PKS~1424+240, fitting the SED with a one-zone model yields a Doppler factor of $\sim100$, while the MOJAVE VLBA data imply a Doppler factor of only $\sim10$ (Aleksi{\'c}, et al.\ 2014a). When the same SED is fit by the two-zone model of Tavecchio et al.\ (2011), the Doppler factor of the fast zone is reduced to $\sim30$, while that of the slower zone has the VLBA-derived value of $\sim10$ (Aleksi{\'c}, et al.\ 2014a). VLBI imaging might detect such two-zone spine-layer jets through observations of transverse emission structures such as limb brightening. We do observe limb brightening in two sources (see $\S$~\ref{transverse}), although for the majority of sources the transverse structure is either unresolved, or it is patchy and complex with multiple emission peaks. However, we note that, because of increased source distance and lower observing frequency, the observations in this paper have about an order of magnitude worse linear resolution compared to the high-frequency observations of limb-brightening in the nearby bright TeV blazars Mrk~421 and Mrk~501 (e.g., Piner et al.\ 2009, 2010). In the spine-layer model for TeV HBLs, the TeV emission comes from the spine while the radio emission comes from the layer, causing different Lorentz factors to be measured in the two spectral bands. However, for TeV radio galaxies, both the gamma-ray and the radio emission can be dominated by the layer (Ghisellini et al.\ 2005), so that consistent Lorentz factors might be expected (no `Doppler Crisis'). VLBI observations of the three known TeV radio galaxies M87 (e.g., Hada et al.\ 2014), Cen~A (e.g., M{\"u}ller et al.\ 2014), and 3C~84 (e.g., Nagai et al.\ 2014) can therefore provide consistency checks on spine-layer models (e.g., Tavecchio \& Ghisellini 2008, 2014). We extended our consideration to the full sample of TeV HBLs in $\S$~\ref{resultsall}, by combining our VLBI data on 20 sources from this paper with other VLBI and gamma-ray data from the literature. Following the approach of Lister et al.\ (2011), we constructed a gamma-ray loudness parameter for the TeV HBLs ($\S$~\ref{loudness}), and found that it spans about two orders of magnitude from extreme gamma-ray loud sources like H~1426+428 to more radio-loud sources like PKS~0301$-$243. There is a significant apparent partial correlation (excluding effects of redshift) between the VLBI and TeV fluxes ($\S$~\ref{correlations}), although Monte Carlo simulations using the method of Pavlidou et al.\ (2012) showed that this correlation may intrinsically be only marginally significant. Such a correlation might suggest that Doppler factors in different jet emission regions are correlated, even if they are significantly different, such as occurs in the model by Lyutikov \& Lister (2010). Note that VLBI core flares correlated with gamma-ray flares in Mrk~421 (Richards et al.\ 2013) also suggest a link between the VLBI emission and the gamma-ray emission in that source. Much more information about the jet kinematics of the TeV HBLs should be revealed through the multi-epoch VLBA monitoring of these 20 sources that is currently underway. At least three additional epochs for each of these sources has been approved on the VLBA, and should be obtained over the next one to two years, in addition to high-frequency imaging of some of the brighter TeV HBLs to investigate transverse jet structures. When added to the 11 TeV HBLs that we have already monitored, and the 7 additional TeV HBLs being monitored by MOJAVE, this will make information on parsec-scale structural changes available for $\sim90\%$ of the currently known TeV HBL population. This is a crucial step toward understanding the jet structure of this group of sources, as the high-energy community looks forward to many more such objects being detected by future TeV telescopes like the Cherenkov Telescope Array (CTA). \vspace*{-0.10in} \acknowledgments The National Radio Astronomy Observatory is a facility of the National Science Foundation operated under cooperative agreement by Associated Universities, Inc. This research has made use of the NASA/IPAC Extragalactic Database (NED) which is operated by the Jet Propulsion Laboratory, California Institute of Technology, under contract with the National Aeronautics and Space Administration. This research has made use of data from the MOJAVE database that is maintained by the MOJAVE team (Lister et al.\, 2009). Part of this research was carried out at the Jet Propulsion Laboratory, Caltech, under a contract with the National Aeronautics and Space Administration. This work was supported by Fermi Guest Investigator grant NNX13AO82G. We also acknowledge helpful comments from the anonymous referee. \vspace*{0.1in} {\it Facilities:} \facility{VLBA ()}
\section{Introduction} Over the past decade, the technique of manipulating ultracold atomic Fermi gases has been well developed and it offers a physical reality to pursue an exotic pairing mechanism, which is referred to as Fulde-Ferrell-Larkin-Ovchinnikov (FFLO) states \cite{Fulde1964,Larkin1964} and has attracted impressive attentions in different physical areas \cite{Casalbuoni2004,Uji2006,Kenzelmann2008,Liao2010,Gerber2014}. In spin-imbalanced Fermi gases, the standard Bardeen-Cooper-Schrieffer (BCS) pairing is not favorable against the FFLO pairing with a finite center-of-mass momentum. Although there is no unambiguous experimental conclusion for the FFLO superfluidity, strong evidence has been seen in a Fermi cloud of $^{6}$Li atoms confined in quasi-one-dimensional harmonic traps near a crossover from a Bose-Einstein condensate (BEC) to a BCS superfluid \cite{Liao2010,Orso2007,Hu2007,Liu2007,Liu2008,Guan2007}. The FFLO pairing is also favored by spin-orbit coupling \cite{YipReview,Barzykin2002,Agterberg2007,Dimitrova2007,Michaeli2012}. Motivated by the recent experimental realization of a synthetic spin-orbit coupling with equal weight combination of Rashba and Dresselhaus components \cite{Lin2011,Wang2012,Cheuk2012,Fu2013}, FF superfluidity - a specific form of the FFLO superfluidity - has been theoretically investigated in spin-orbit coupled atomic Fermi gases \cite{reviewSOC,Dong2013,Zheng2013,Wu2013,Liu2013a,Dong2013NJP,Zhou2013,Iskin2013,Shenoy2013}. In the case of a Rasbha spin-orbit coupling, topological superfluidity is argued to be achievable \cite{Zhang2008,Sato2009,Sau2010,Oreg2010,Zhu2011,Liu2012a,Liu2012b,Wei2012}, although the underlying pairing is of \textit{s}-wave character. It turns out that the topological superfluidity and FF superfluidity are compatible. As a result, novel topological FF superfluids have also been proposed \cite{Chen2013,Liu2013b,Qu2013,Zhang2013,Cao2014,Hu2014,Jiang2014}. In particular, in a recent Letter, some of us have predicted that a \emph{gapless} topological FF superfluid may appear in a two-dimensional (2D) spin-orbit coupled atomic Fermi gas with both in-plane and out-of-plane Zeeman fields \cite{Cao2014}. The purpose of the present work is to provide more details about such an interesting superfluid phase and to discuss its thermodynamic stability by considering the superfluid density and superfluid transition temperature. It is well known that at finite temperatures the superfluidity of 2D atomic Fermi gases is characterized by the vortex-antivortex (V-AV) binding. The relevant mechanism is the Berezinskii-Kosterlitz-Thouless (BKT) transition occurring at a characteristic temperature $T_{\mathrm{{BKT}}}$ \cite{Berezinskii1971,Kosterlitz1972}. Below the critical BKT temperature, a V-AV binding state has a lower free energy and hence superfluidity emerges. The BKT transition was theoretically investigated long time ago in a 2D fermionic system without spin-orbit coupling \cite{Randeria1989,Gusynin1999,Loktev2001}. Following the recent experimental advances, there have been several theoretical investigations about the superfluid density and critical BKT temperature in 2D spin-orbit coupled Fermi gases with BCS pairing \cite{He2012,Gong2012,He2013}. In the case of a large out-of-plane Zeeman field, the temperature region for experimentally observing topological BCS superfluids and related Majorana fermions has been discussed \cite{Gong2012,He2013}. However, the BKT physics of a spin-orbit coupled FF superfluid - which can be either gapped or gapless, topologically trivial or non-trivial - has so far not been addressed. In this work, we explore this interesting issue and study the superfluid density tensor and BKT transition of a 2D Rasbha spin-orbit coupled Fermi gas in the presence of both in-plane and out-of-plane Zeeman fields. By calculating the superfluid density tensor, we obtain the superfluid phase stiffness as functions of the temperature, spin-orbit coupling strength, binding energy (that characterizes the interatomic interaction strength), in-plane and out-of-plane Zeeman fields. This allows us to determine the critical BKT temperature of the system in four different FF superfluid phases \cite{Cao2014}, with a given set of parameters. Our main results may be summarized as follows: (i) At zero temperature with an applied in-plane Zeeman field in the \textit{x}-direction, the component $n_{s,xx}$ of the superfluid density tensor always changes discontinuously when the system continuously evolves from a gapped FF into a gapless FF phase. The component $n_{s,yy}$ is larger than $n_{s,xx}$ except for a narrow parameter space where the FF momentum is sufficiently large. The two components of the superfluid density tensor decrease monotonically as the temperature increases. (ii) All the four FF superfluid phases have significant critical BKT temperature, except for the parameter region with very small spin-orbit coupling and/or binding energy, or with very large in-plane and/or out-of-plane Zeeman fields. The critical BKT temperature can be enhanced by increasing the binding energy. But it does not increase monotonically as the spin-orbit coupling strength increases. The rest of the paper is organized as follows. In the next section, we briefly describe the mean-field theoretical framework, and clarify the BKT physics in two dimensions and the related Kosterlitz-Thouless-Nelson (KT-Nelson) criterion for phase transition. Then, we present the expressions for the superfluid density tensor and superfluid phase stiffness. The critical BKT temperature is determined by applying the KT-Nelson criterion. In Sec. III, we first present the finite-temperature phase diagram of the system and then discuss in detail the results on the superfluid density tensor and critical BKT temperature. Finally, Sec. IV is devoted to the conclusions and outlooks. \section{Model Hamiltonian and mean-field theory} We start by considering a 2D spin-orbit coupled two-component Fermi gas near a broad Feshbach resonance with the Rashba spin-orbit coupling $\lambda\bm{\hat{\sigma}}\cdot\bm{\hat{\mathrm{k}}}$, in-plane ($h_{x}$) and out-of-plane ($h_{z}$) Zeeman fields \cite{notation}. The system can be well described by the following single-channel Hamiltonian, \begin{equation} \text{\ensuremath{\mathcal{H}}}=\int d{\bf r}\left[\mathcal{H}_{0}+\mathcal{H}_{int}\right],\label{eq:totHami} \end{equation} where \begin{equation} {\cal H}_{0}=\psi^{\dagger}(\bm{r})\left(\hat{\xi}_{{\bf k}}+\lambda\bm{\hat{\sigma}}\cdot\bm{\hat{\mathrm{k}}}-h_{z}\hat{\sigma}_{z}-h_{x}\hat{\sigma}_{x}\right)\psi(\bm{r})\label{eq:spHami} \end{equation} is the single-particle Hamiltonian and \begin{equation} \text{\ensuremath{\mathcal{H}}}_{int}=U_{0}\psi_{\uparrow}^{\dagger}({\bf r})\psi_{\downarrow}^{\dagger}({\bf r})\psi_{\downarrow}({\bf r})\psi_{\uparrow}({\bf r}) \end{equation} is the density of interaction Hamiltonian in which the bare interaction strength $U_{0}$ is to be regularized as \begin{equation} \frac{1}{U_{0}}=-\frac{1}{\mathcal{S}}\sum_{\mathbf{k}}\frac{1}{\hbar^{2}\mathbf{k}^{2}/m+E_{b}}, \end{equation} with $\mathcal{S}$ being the area of the system and $E_{b}$ the two-particle binding energy that physically characterizes the interaction strength. In the single-particle Hamiltonian, $\lambda$ is the Rashba spin-orbit coupling strength and we have used the following notations: (1) $\hat{\xi}_{{\bf k}}\equiv-\hbar^{2}\nabla^{2}/(2m)-\mu$ with the atomic mass $m$ and chemical potential $\mu$; (2) $\bm{\mathrm{\hat{k}}}=(\hat{k}_{x},\hat{k}_{y})$, where $\hat{k}_{x}=-i\partial_{x}$ and $\hat{k}_{y}=-i\partial_{y}$ are momentum operators; and (3) $\bm{\hat{\sigma}}=(\hat{\sigma}_{x},\hat{\sigma}_{y})$, the Pauli matrices. We have also used $\psi({\bf r})=[\psi_{\uparrow}(\bm{r}),\psi_{\downarrow}(\bm{r})]^{T}$ ($\psi^{\dagger}(\bm{r})=[\psi_{\uparrow}^{\dagger}(\bm{r}),\psi_{\downarrow}^{\dagger}(\bm{r})]$) to collectively denote the fermion field operator for creating (annihilating) an atom at ${\bf r}$ with a specific spin $\sigma=\uparrow,\downarrow$. \subsection{Mean-field theory} We solve the model Hamiltonian Eq. (\ref{eq:totHami}) by using the functional path-integral approach \cite{reviewSOC,He2013,Hu2011,Jiang2011}. At the inverse finite temperature $\beta=1/(k_{B}T)$, the partition function can be written as, \begin{equation} \text{\ensuremath{\mathcal{Z}}}=\int\mathcal{\mathcal{D}}\psi\left(\mathbf{r},\tau\right)\mathcal{D}\bar{\psi}\left(\mathbf{r},\tau\right)\exp\left\{ -\mathcal{A\left[\psi,\bar{\psi}\right]}\right\} ,\label{eq:partition1} \end{equation} where \begin{equation} \mathcal{A}\left[\psi,\bar{\psi}\right]=\int_{0}^{\beta}d\tau\int d\bm{r}\bar{\psi}\partial_{\tau}\psi+\int_{0}^{\beta}d\tau\mathcal{H}\left(\psi,\bar{\psi}\right).\label{eq:action} \end{equation} Here, the field operators $\psi$ and $\psi^{\dagger}$ in the model Hamiltonian $\mathcal{H}$ have been replaced with the corresponding Grassmann variables $\psi(\mathbf{r},\tau)$ and $\bar{\psi}(\mathbf{r},\tau)$, respectively. Following the standard procedure \cite{reviewSOC}, the interaction term in the Hamiltonian is decoupled using the Hubbard-Stratonovich transformation. Introducing the auxiliary complex pairing field $\phi(\mathbf{r},\tau)=-U_{0}\psi_{\downarrow}(\mathbf{r},\tau)\psi_{\uparrow}(\mathbf{r},\tau)$, and integrating out the Grassmann fields, the partition function becomes \begin{equation} \text{\ensuremath{\mathcal{Z}}}=\int\mathcal{\mathcal{D}\phi}\left(\mathbf{r},\tau\right)\mathcal{D}\bar{\phi}\left(\mathbf{r},\tau\right)\exp\left\{ -\mathcal{A}_{eff}\left[\phi,\bar{\phi}\right]\right\} ,\label{eq:partition2} \end{equation} where in the saddle-point approximation (i.e., mean-field treatment by replacing $\phi(\mathbf{r},\tau)$ with a static pairing field $\Delta(\mathbf{r})$), the effective action $\text{\ensuremath{\mathcal{A}}}_{eff}$ takes the form, \begin{equation} \text{\ensuremath{\mathcal{A}}}_{mf}=\beta\sum_{\mathbf{k}}\hat{\xi}_{\mathbf{k}}-\int_{0}^{\beta}d\tau\int d\bm{r}\frac{\left|\Delta\right|{}^{2}}{U_{0}}-\frac{1}{2}\textrm{Tr}\ln\left[-G^{-1}\right].\label{eq:effaction} \end{equation} In the above expression, $G^{-1}\left(\mathbf{r},\tau\right)=-\partial_{\tau}-\mathcal{H}_{BdG}$ is the inverse single-particle Green function in the Nambu-Gorkov representation, with a mean-field Bogoliubov Hamiltonian, \begin{equation} \mathcal{H}_{BdG}=\left[\begin{array}{cc} H_{0}(\mathbf{\hat{k}}) & -i\Delta(\mathbf{r})\hat{\sigma}_{y}\\ i\Delta(\mathbf{r})\hat{\sigma}_{y} & -H_{0}^{*}\left(-\mathbf{\hat{k}}\right) \end{array}\right], \end{equation} where $H_{0}\equiv\hat{\xi}_{{\bf k}}+\lambda\bm{\hat{\sigma}}\cdot\bm{\hat{\mathrm{k}}}-h_{z}\hat{\sigma}_{z}-h_{x}\hat{\sigma}_{x}$. In the presence of the in-plane Zeeman field $h_{x}$, it is known that the pairing field takes the FF form $\Delta(\mathbf{r})=\Delta e^{iQx}$, with a finite center-of-mass momentum of the pairs $\mathbf{Q}=Q\mathbf{e}_{x}$ \cite{Zheng2013,Wu2013,Liu2013a,Dong2013NJP,Zhou2013}. This helical phase was earlier studied in the context of noncentrosymmetric superconductors \cite{Agterberg2007,Dimitrova2007}. The resulting mean-field thermodynamic potential $\Omega_{mf}=k_{B}T\mathcal{A}_{mf}$ reads, \begin{equation} \Omega_{mf}=\sum_{\mathbf{k}}\hat{\xi}_{\mathbf{k}}-\mathcal{S}\frac{\Delta^{2}}{U_{0}}-\frac{k_{B}T}{2}\sum_{\mathbf{k},i\omega_{m}}\ln\det\left[-G^{-1}\left(\mathbf{k},i\omega_{m}\right)\right],\label{eq:thermaldynamic} \end{equation} where $G^{-1}(\mathbf{k},i\omega_{m})$ is the inverse Green function in momentum space and $\omega_{m}=\pi(2m+1)/\beta$ with integer $m$ is the fermionic Matsubara frequency. Making use of the inherent particle-hole symmetry of the BdG Hamiltonian, we find that, \begin{equation} \det\left[-G^{-1}\left(\mathbf{k},i\omega_{m}\right)\right]=\prod_{\eta=1,2}\left[\left(i\omega_{n}\right){}^{2}-\left(E_{\bm{\mathrm{k}}\eta}^{\nu=+}\right){}^{2}\right], \end{equation} where $E_{\mathrm{\bm{k}\eta}}^{\nu}$ is the quasi-particle energy, obtained by diagonalizing $\mathcal{H}_{BdG}$ with the FF pairing field $\Delta(\mathbf{r})=\Delta e^{iQx}$ \cite{Liu2013a,Dong2013NJP}. The superscript $\nu\in(+,-)$ represents the particle ($+$) or hole ($-$) branch and the subscript $\text{\ensuremath{\eta\in}(1,2)}$ denotes the upper ($1$) or lower ($2$) branch split by the spin-orbit coupling \cite{Liu2013a,Hu2011,Jiang2011}. By summing over the Matsubara frequency, the mean-field thermodynamic potential takes the form, \begin{eqnarray} \Omega_{mf} & = & \frac{1}{2}\sum_{\mathbf{\bm{k}}}\left(\xi_{\mathbf{\bm{k}}+\mathbf{\bm{Q}}/2}+\xi_{\mathbf{\bm{k}}-\mathbf{\bm{Q}}/2}\right)-\frac{1}{2}\sum_{\mathbf{\bm{k}\eta}}|E_{\bm{\mathrm{\bm{k}}}\eta}^{+}|\nonumber \\ & & -k_{B}T\sum_{\mathbf{k\eta}}\ln\left(1+e^{-|E_{\mathbf{k}\eta}^{+}|/k_{B}T}\right)-\mathcal{S}\frac{\Delta^{2}}{U_{0}}. \end{eqnarray} Here the term $\sum_{\mathbf{k}}\hat{\xi}_{\mathbf{k}}$ is replaced by $(1/2)\sum_{\mathbf{k}}(\xi_{\mathbf{k}+\mathbf{Q}/2}+\xi_{\mathbf{k}-\mathbf{Q}/2})$, in order to cancel the leading divergence of the term $(1/2)\sum_{\mathbf{\bm{k}\eta}}|E_{\bm{\mathrm{\bm{k}}}\eta}^{+}|$. For a given set of parameters, for example, the temperature $T$, binding energy $E_{b}$ etc., different superfluid phases can be determined using the self-consistent stationary conditions: \begin{eqnarray} \frac{\partial\Omega_{mf}}{\partial\Delta} & = & 0,\\ \frac{\partial\Omega_{mf}}{\partial Q} & = & 0, \end{eqnarray} as well as the conservation of total atom number, \begin{equation} n=-\frac{1}{\mathcal{S}}\frac{\partial\Omega_{mf}}{\partial\mu}, \end{equation} where $n=N/\mathcal{S}$ is the number density. At a given temperature, the ground state has the lowest free energy $F=\Omega_{mf}+\mu N$. \subsection{Superfluid density tensor} An important quantity to characterize the anisotropic superfluid properties of a 2D spin-orbit coupled Fermi gas is the superfluid density tensor. In the case of BCS pairing, the superfluid density tensor may be analytically derived within mean-field framework \cite{He2012,Gong2012,Zhou2012}, yet the formalism has not been obtained for a FF superfluid. According to the definition of the superfluid density, we calculate it by applying a phase twist to the order parameter, $\Delta_{twist}(\mathbf{v}_{s})=\Delta(\mathbf{r})e^{i\mathbf{q}\cdot\bm{r}}$ , which boosts the system with a uniform superfluid flow at a velocity $\mathbf{v}_{s}=\hbar\mathbf{q}/2m$ \cite{Zhou2012,Taylor2006,Fukushima2007}. Here $\Delta(\mathbf{r})$ is the equilibrium FF order parameter. Physically, only the superfluid component moves under the influence of the superfluid flow. Thus, as the result of this boost, the thermodynamic potential assumes the following form in the limit of small velocity, \begin{equation} \Omega\left(\mathbf{v}_{s}\right)\simeq\Omega\left(\mathbf{v}_{s}=0\right)+\frac{1}{2}m\mathcal{S}\sum_{ij}n_{s,ij}v_{si}v_{sj}, \end{equation} where $n_{s,ij}$ ($i,j=x,y$) is the superfluid density tensor. Therefore, we immediately obtain \cite{Zhou2012,Taylor2006,Fukushima2007}, \begin{equation} n_{s,ij}=\frac{1}{\mathcal{S}}\frac{4m}{\hbar^{2}}\left[\frac{\partial^{2}\Omega\left(\mathbf{v}_{s}\right)}{\partial q_{i}\partial q_{j}}\right]_{\mathbf{q}=0},\label{eq:ns} \end{equation} where $\Omega\left(\mathbf{v}_{s}\right)$ should be calculated with $\Delta_{twist}(\mathbf{v}_{s})$ in the presence of the phase twist. The above relation for the superfluid density tensor is rigorous. In this work, consistent with the mean-field treatment for thermodynamics, in Eq. (\ref{eq:ns}) we shall approximate the thermodynamic potential $\Omega\left(\mathbf{v}_{s}\right)$ by its mean-field value $\Omega_{mf}\left(\mathbf{v}_{s}\right)$. \subsection{The KT-Nelson criterion for $T_{BKT}$} The BKT transition in 2D is peculiar, associated with the spontaneous vortex formation. A unique feature of such a transition is a universal jump in the superfluid density (tensor), characterized by the KT-Nelson criterion for the critical BKT temperature \cite{Nelson1977}. It may be explained by using the following simple physical picture for the spontaneous creation of a single vortex at finite temperature $T$. In the absence of spin-orbit coupling and Zeeman fields, let us consider an \emph{isotropic} Fermi superfluid in a circular disk geometry, with a radius of $R\rightarrow\infty$. The kinetic energy cost for creating a single vortex at the origin $\mathbf{r}=\mathbf{0}$ is simply given by, \begin{equation} E_{V}\simeq\frac{1}{2}mn_{s}\int_{\xi}^{R}d^{2}\mathbf{r}\left(\frac{\hbar}{2mr}\right)^{2}=\frac{\hbar^{2}\pi}{4m}n_{s}\ln\left(\frac{R}{\xi}\right), \end{equation} where $\xi$ is the size of the vortex core. The associated entropy can be calculated by the number of distinct positions at which the vortex can be placed, \begin{equation} S_{V}\simeq k_{B}\ln\left(\frac{\pi R^{2}}{\pi\xi^{2}}\right)=2k_{B}\ln\left(\frac{R}{\xi}\right). \end{equation} From these two expressions, we see that the free energy associated with the formation of a single vortex is, \begin{equation} F_{V}=E_{V}-TS_{V}\simeq2\left(\frac{\pi}{2}\frac{\hbar^{2}}{4m}n_{s}-k_{B}T\right)\ln\left(\frac{R}{\xi}\right). \end{equation} It is clear that the free energy changes its sign at a characteristic temperature $T_{BKT}$ determined by \begin{equation} k_{B}T_{BKT}=\frac{\pi}{2}\mathcal{J},\label{eq:BKT} \end{equation} where $\mathcal{J}=\hbar^{2}n_{s}/(4m)$ is the superfluid phase stiffness. This is the well-known KT-Nelson criterion \cite{Nelson1977}. As $\ln(R/\xi)$ diverges in the thermodynamic limit $R\rightarrow\infty$, the temperature $T_{BKT}$ separates two qualitatively different regimes. At $T>T_{BKT}$, the free energy is very large and negative, suggesting the spontaneous creation of a free vortex with either positive or negative circulation. While at $T<T_{BKT}$, vortices with opposite circulation will bind together and generate coherence. The spontaneous creation of free vortex suggests that the loss of the phase coherence of the system occurs suddenly. It leads to a universal jump in the superfluid phase stiffness or superfluid density, as can be seen clearly from the KT-Nelson criterion, Eq. (\ref{eq:BKT}). In the case of an \emph{anisotropic} superfluid, we need to define a superfluid density tensor \begin{equation} \mathscr{\mathcal{N}}_{s}=\left[\begin{array}{cc} n_{s,xx} & n_{s,xy}\\ n_{s,yx} & n_{s,yy} \end{array}\right]. \end{equation} The associated superfluid phase stiffness takes the form, \begin{equation} \mathcal{J}=\frac{\hbar^{2}}{4m}\left(\det\mathscr{\mathcal{N}}_{s}\right)^{1/2}=\frac{\hbar^{2}}{4m}\sqrt{n_{s,xx}n_{s,yy}}, \end{equation} where in the last equation, we use the fact that $n_{s,xy}=n_{s,yx}=0$, which holds for the system considered in this work. It is worth noting that although Eq. (\ref{eq:BKT}) is obtained by drawing a simple physical picture, it would be a rigorous criterion for the BKT transition. Indeed, the KT-Nelson criterion was first obtained by using a renormalization group analysis \cite{Nelson1977}. For a microscopic derivation, we may consider the contribution of the pair fluctuations around the saddle-point solution $\delta\phi\left(\mathbf{q},i\nu_{n}\right)$ to the action $\delta\mathcal{A}$, which, at the \emph{Gaussian} (quadratic) level, is given by \cite{He2012,He2013,Salasnich2013,Yin2014}, \begin{equation} \delta\mathcal{A}=\frac{1}{2}\sum_{\mathcal{Q}=\mathbf{q},i\nu_{n}}\left[\delta\phi^{\dagger}\left(\mathcal{Q}\right),\delta\phi\left(-\mathcal{Q}\right)\right]\mathbf{M}\left[\begin{array}{c} \delta\phi\left(\mathcal{Q}\right)\\ \delta\phi^{\dagger}\left(-\mathcal{Q}\right) \end{array}\right], \end{equation} where the $2\times2$ matrix \begin{equation} \mathbf{M}\equiv\left[\begin{array}{cc} M_{11}\left(\mathcal{Q}\right), & M_{12}\left(\mathcal{Q}\right)\\ M_{21}\left(\mathcal{Q}\right), & M_{22}\left(\mathcal{Q}\right) \end{array}\right] \end{equation} is the inverse two-particle (pair) propagator and its elements can be evaluated with the mean-field fermionic Green function $G(\mathbf{k},i\omega_{m})$. In the case of BCS pairing without the in-plane Zeeman field, the expression of the inverse pair propagator $\bm{\mathrm{M}}$ can be analytically obtained \cite{He2013,Salasnich2013}. In particular, in the limit of long wavelength, the matrix elements of $\bm{\mathrm{M}}$ can be expanded as functions of small $\bm{\mathrm{k}}$ and $\omega$. By separating the phase fluctuation and amplitude (density) fluctuation, the low-energy physics of the system can be found to be governed by the well-known classical spin XY model \cite{He2013,Salasnich2013}, which is the prototype of the BKT physics. In this way, one microscopically derives the superfluid phase stiffness $\mathcal{J}$ and the KT-Nelson relation. The resulting expression for the superfluid phase stiffness \emph{coincides} with the mean-field phase stiffness obtained, for example, by using the mean-field thermodynamic potential in Eq. (\ref{eq:ns}). In our FF case, the expression of the superfluid phase stiffness could be derived in a similar manner. However, in this case, the analytical expression of the inverse pair propagator $\bm{\mathrm{M}}$ is more difficult to obtain, although we can numerically sum over the bosonic Matsubara frequency $i\nu_{n}$. Therefore, to calculate the superfluid phase stiffness, we prefer to directly use Eq. (17) with a mean-field thermodynamic potential. \subsection{Pair fluctuations beyond mean-field} To close this section, we briefly discuss how to improve the mean-field theory. An immediate idea is to work out the Gaussian correction to the action, $\delta\mathcal{A}$, and then use the improved thermodynamic potential around the saddle point $\Delta(\mathbf{r})=\Delta e^{iQx}$ \cite{Taylor2006,Fukushima2007,Hu2006}, \begin{equation} \Omega_{GPF}=\Omega_{mf}+k_{B}T\sum_{\mathcal{Q}=\mathbf{q},i\nu_{n}}\ln\mathbf{M}\left(\mathcal{Q}\right), \end{equation} to calculate the equation of state through the standard thermodynamic relations and the superfluid density tensor via Eq. (\ref{eq:ns}). In this way, the thermodynamics and the superfluid density tensor of the system can be consistently determined at the same level of approximation. Alternatively, we may also consider using $\Omega_{GPF}$ to determine the chemical potential $\mu$ and then calculate the superfluid density tensor using the mean-field expression. However, as the trade-off of this cheap treatment, we may have an inconsistency. The resulting critical BKT temperature could be less reliable. For a detailed discussion, we refer to the recent work by Tempere and Klimin \cite{Tempere2014}. \section{Results and discussions} Using the above-mentioned mean-field theoretical framework, we have systematically explored the low-temperature phase diagram and the thermodynamic stability of different exotic Fulde-Ferrell superfluid phases. In our numerical calculations, we take the Fermi wavevector $k_{F}=\sqrt{2\pi n}$ and the Fermi energy $E_{F}=\hbar^{2}k_{F}^{2}/(2m)$ as the units for wavevector and energy, respectively. For a typical set of parameters (i.e., default parameters), we use the interaction parameter $E_{b}=0.2E_{F}$, spin-orbit coupling strength $\lambda=E_{F}/k_{F}$, in-plane Zeeman field $h_{x}=0.4E_{F}$, out-of-plane Zeeman field $h_{z}=0.1E_{F}$ and temperature $T=0.05T_{F}$. \subsection{Low-temperature phase diagrams} In the recent Letter \cite{Cao2014}, we have discussed the phase diagram and the appearance of an interesting gapless topological Fulde-Ferrell superfluid at a weak interaction strength parameterized by $E_{b}=0.2E_{F}.$ Experimentally, it is most likely that the measurement will be carried out at a stronger interaction strength, where the superfluid transition temperature is anticipated to be higher. In order to optimize the experimental condition for observing the gapless topological superfluid, here we present a systematic study with varying binding energy, from the weakly interacting BCS side to the strongly interacting BEC-BCS crossover regime. \begin{figure} \begin{centering} \includegraphics[clip,width=0.48\textwidth]{fig1_phasediagram} \par\end{centering} \protect\caption{(color online) Phase diagrams of a 2D spin-orbit coupled atomic Fermi gas at a broad Feshbach resonance and at a typical low temperature $0.05T_{F}$ with (a) $h_{z}=0.1E_{F}$ or (b) $h_{x}=0.4E_{F}$. The strength of spin-orbit coupling is $\lambda=E_{F}/k_{F}$. There are four superfluid phases: gFF, nFF, tnFF and tgFF (whose phase stiffness $\pi\mathcal{J}/2$ - in units of $E_{F}$ - is illustrated in color), as well as a pseudogap phase (grey area). We treat the system as a normal gas (shown in white) when the pairing gap $\Delta<10^{-3}$. In the gapless topological phase, the notations tnFF$_{1}$, tnFF$_{2}$ and tnFF$_{3}$ distinguish different zero-energy contours in energy spectrum. For details, see the contour plots in Fig. \ref{fig2}.} \label{fig1} \end{figure} In Fig. 1, we report two phase diagrams at the typical low temperature $T=0.05T_{F}$ on the plane of $E_{b}$-$h_{x}$ (a) or $E_{b}$-$h_{z}$ (b). The superfluid phase stiffness $\pi\mathcal{J}/2$ in different phases is color illustrated and its detailed behavior will be discussed in the next subsection. The superfluid phases are determined using the KT-Nelson criterion $\pi\mathcal{J}(T=0.05T_{F})/2>k_{B}T=0.05E_{F}$. Obviously, there is a pseudogap regime (shown in grey), in which the pairing order parameter is finite but the superfluid phase stiffness is not large enough to drive the BKT transition. A better understanding of the pseudogap phase requires a careful treatment of strong phase fluctuations. It is out of the scope of the present paper. \subsubsection{gapless topological transition} It is known from previous studies \cite{Qu2013,Zhang2013,Cao2014} that the combined effect of spin-orbit coupling, in-plane and out-of-plane Zeeman fields may induce several exotic superfluid phases: gapped FF (gFF), gapless FF (nFF), gapless topological FF (tnFF) and gapped topological FF (tgFF), classified by considering whether the system has a bulk-gapped and/or topologically non-trivial energy spectrum. In the literature, the topological superfluidity was firstly studied with an out-of-plane Zeeman field only \cite{reviewSOC,Zhang2008}. In that case, topological phase transition can be driven by increasing the out-of-plane Zeeman field $h_{z}$ above a threshold \begin{equation} h_{z,c}=\sqrt{\Delta^{2}+\mu^{2}},\label{eq:tpt} \end{equation} at which the dispersions of the particle- and hole-branches touch each other at the single point $\bm{\mathrm{\bm{k}}}=0$, meanwhile the bulk excitation gap closes. Afterwards, the topology of the Fermi surface dramatically changes and the excitation gap re-opens \cite{reviewSOC,Hasan2010,Qi2011}. It is straightforward to understand the single-point closure of the excitation gap, since the Fermi surface is always rotationally symmetric. This also implies that the resulting topological superfluid must be gapped in the bulk. However, such a scenario may be greatly altered by the presence of a non-zero in-plane Zeeman field, which favors the FF pairing with a finite center-of-mass momentum and consequently breaks the rotational symmetry of the Fermi surface. In the case of a small in-plane Zeeman field, the rotational symmetry breaking of the energy spectrum is not significant. Although the system becomes a FF superfluid, its bulk excitation gap still closes at the single point $\mathrm{\bm{\mathrm{k}}=0}$, accompanied by the change of the topology of the Fermi surface. An example is the transition from gFF to tgFF shown in Fig. \ref{fig1}(b) at large binding energy $E_{b}>0.3E_{F}$, where the in-plane Zeeman field is effectively weak. As a result, the picture of the out-of-plane field induced topological phase transition remains unchanged \cite{Liu2013b,Qu2013,Zhang2013}. When the in-plane Zeeman field keeps increasing over a threshold $h_{x,c1}$, however, the closure of the excitation gap and the change of the topology of the Fermi surface may not occur at the same time. A gapless superfluid phase - referred to as nFF - may emerge in the first place at $\bm{\mathrm{\bm{k}}}\neq0$. The nodal points with $E_{\eta=2}^{\nu}(k_{x},k_{y})=0$ form two disjoint loops (see, for example, the transition from gFF to nFF in Fig. \ref{fig1}(a)). The topology of the Fermi surface only changes when the in-plane Zeeman field further increases up to another critical value $h_{x,c2}$, at which the two nodal loops connect at $\text{\ensuremath{\bm{\mathrm{k}}}}=0$. We refer to the previous work Ref. \cite{Cao2014} for a detailed characterization of the gapless topological transition. \subsubsection{Binding energy dependence of the phase diagram} It can now be understood that both the in-plane and out-of-plane fields can drive the topological phase transition, but the underlying property of the resulting topological phase, in terms of the gapless or gapped bulk spectrum, depends critically on the relative strength of the two fields. The gapless topological FF superfluid (tnFF) intentionally emerges in the parameter regime where $h_{x}$ is larger enough relative to $h_{z}$. This is particularly clear from Fig. \ref{fig1}(a), where we have fixed the strength of the out-of-plane Zeeman field to $h_{z}=0.1E_{F}$. The tnFF phase accounts for most of the space for topological phases. It is remarkable that the window of the tnFF superfluid remains very significant when the binding energy increases up to $0.5E_{F}$, suggesting the use of a large interaction strength near Feshbach resonances, for the purpose of having a larger BKT transition temperature to observe the exotic tnFF phase. On the contrary, Fig. \ref{fig1}(b) - where we have fixed the in-plane Zeeman field to $h_{x}=0.4E_{F}$ - clearly reveals that the gapped topological FF superfluid (tgFF) occupies most of the space for topological phases, when the out-of-plane Zeeman field is larger than the in-plane Zeeman field. In this case, the tnFF phase is restricted to the parameter space with a small out-of-plane Zeeman field and a weak interaction strength, as one may anticipate. \subsubsection{Different gapless topological superfluid phases} \begin{figure} \begin{centering} \includegraphics[clip,width=0.48\textwidth]{fig2_spectrum} \par\end{centering} \protect\caption{(color online) Dispersion relation of the lower branch $E_{\eta=2}^{\nu}(k_{x},k_{y}=0)$ (left panel, red curves for particle excitations $\nu=+$ and blue curves for hole excitations $\nu=-$) and the corresponding contour of zero-energy nodes (right panel). (a) and (b) correspond to the red point in Fig. \ref{fig1} for the tnFF$_{1}$ phase, (c) and (d) the yellow point for the tnFF$_{2}$ phase and, (c) and (f) the magenta point for the tnFF$_{3}$ phase.} \label{fig2} \end{figure} It is interesting that the gapless topological FF superfluid may be further classified into different categories (tnFF$_{1}$, tnFF$_{2}$ and tnFF$_{3}$), according to the number and position of its disjoint loops of nodal points, as shown in the right panel of Fig. \ref{fig2}. The tnFF$_{1}$ superfluid is most common and has two nodal loops, one for the particle branch (red loop) and another for the hole branch (blue loop). The tnFF$_{3}$ superfluid also has two nodal loops. However, the loops for the particle and hole branches exchange their position. It occurs only at large in-plane Zeeman field and binding energy. The tnFF$_{2}$ seems to connect the tnFF$_{1}$ and tnFF$_{3}$ phases. It has four disjoint nodal loops and exists only in a very narrow parameter space (see, for example, Fig. \ref{fig1}(a)). We note that the two gapless topological phases, tnFF$_{1}$ and tnFF$_{3}$, may also be intervened by a \emph{gapped} topological phase, in which there is no nodal loop at all. \subsection{Superfluid density} Having determined the low-temperature phase diagram, we are in position to understand the superfluid density and the critical BKT temperature of different superfluid phases, which have been only briefly mentioned in our previous Letter \cite{Cao2014}. In the presence of spin-orbit coupling, it is known that the superfluid density is a tensor \cite{He2012,He2013}. We then have to consider both diagonal elements of the superfluid density tensor, $n_{s,xx}$ and $n_{s,yy}$. \begin{figure} \begin{centering} \includegraphics[clip,width=0.48\textwidth]{fig3_superfluiddensity_hxhz} \par\end{centering} \protect\caption{(color online) Diagonal elements of the superfluid density tensor as a function of $h_{x}$ and $h_{z}$ at zero temperature (left panel) and at a finite temperature $T=0.05T_{F}$ (right panel). The superfluid density is measured in units of the total density $n=k_{F}^{2}/(2\pi)$. In (a) and (b), the out-of-plane Zeeman field strength $h_{z}=0.1E_{F}$. In (c) and (d), the in-plane Zeeman field strength $h_{x}=0.4E_{F}$. Other parameters are $E_{b}=0.2E_{F}$ and $\lambda=E_{F}/k_{F}$.} \label{fig3} \end{figure} In Fig. \ref{fig3}, we present the Zeeman field dependence of $n_{s,xx}$ and $n_{s,yy}$ at zero temperature (left panel, a and c) and at a finite temperature $T=0.05T_{F}$ (right panel, b and d). In general, as a consequence of the in-plane Zeeman field applied along the $x$-axis, $n_{s,xx}$ is smaller than $n_{s,yy}$, except at extremely low temperature and sufficiently large Zeeman fields. At zero temperature, $n_{s,xx}$ initially decreases with increasing Zeeman fields and exhibits a sudden drop when the system evolves from the gFF phase into the nFF phase at the threshold $h_{x,c1}$ (or $h_{z,c1}$). At $h_{x}>h_{x,c1}$ (or $h_{z}>h_{z,c1}$) it then rises up gradually and is always enhanced by the Zeeman field. Apart from the discontinuous jump, similar Zeeman-field dependence of the superfluid density has been reported for a gapped BCS topological superfluid across the topological phase transition \cite{He2013}. Compared with the non-monotonic field dependence of $n_{s,xx}$, we always find that $n_{s,yy}$ decreases continuously with increasing the Zeeman field. Instead of the sudden drop, a kink is observed at the transition from the gFF phase to the nFF phase. The behavior of the superfluid density is profoundly affected by a nonzero temperature. Already at $T=0.05T_{F}$, the discontinuous drop in $n_{s,xx}$ is smoothed out, leaving a broad minimum with a width $\Delta h_{x,z}\sim2k_{B}T=0.1E_{F}$. Moreover, at the large Zeeman field $h_{x,z}\sim0.6E_{F}$, $n_{s,xx}$ starts to decrease with increasing the Zeeman field. At even higher temperature (not shown in the figure), the local minimum in $n_{s,xx}$ may disappear. \begin{figure} \begin{centering} \includegraphics[clip,width=0.48\textwidth]{fig4_superfluiddensity_temperature} \par\end{centering} \protect\caption{(color online) Temperature dependence of the diagonal elements of the superfluid density tensor, at the six points shown in Fig. 1(a): (a) $E_{b}=0.21E_{F}$ and $h_{x}=0.58E_{F}$, the tnFF$_{1}$ phase; (b) $E_{b}=0.21E_{F}$ and $h_{x}=0.71E_{F}$, the tnFF$_{2}$ phase; (c) $E_{b}=0.33E_{F}$ and $h_{x}=0.8E_{F}$, the tnFF$_{3}$ phase; (d) $E_{b}=0.4E_{F}$ and $h_{x}=0.789E_{F}$, the tgFF phase; (e) $E_{b}=0.21E_{F}$ and $h_{x}=0.2E_{F}$, the gFF phase; and (f) $E_{b}=0.1E_{F}$ and $h_{x}=0.33E_{F}$, the nFF phase. The superfluid density is measured in units of the total density $n=k_{F}^{2}/(2\pi)$. Other parameters are $h_{z}=0.1E_{F}$ and $\lambda=E_{F}/k_{F}$.} \label{fig4} \end{figure} In Fig. \ref{fig4}, we report the temperature dependence of the superfluid density at six typical sets of parameters, which correspond to different superfluid phases at $T=0.05T_{F}$, as shown in Fig. 1(a). $n_{s,xx}$ and $n_{s,yy}$ decrease as temperature increases, in agreement with the common idea that the superfluid component should be gradually destroyed by thermal excitations. It is remarkable that for the gapless topological tnFF$_{1}$ phase (see Fig. \ref{fig4}(a)), the superfluid density does not decrease rapidly with increasing temperature, implying a sizable critical BKT transition temperature for its experimental observation, as we shall discuss in greater detail in the next subsection. In contrast, the superfluid density of other two gapless topological phases (tnFF$_{2}$ and tnFF$_{3}$ in Figs. \ref{fig4}(b) and \ref{fig4}(c), respectively) is more sensitive to temperature and vanishes at $T\sim0.1T_{F}$, probably due to their large Zeeman fields. \subsection{Critical BKT temperature and finite-temperature phase diagrams} \begin{figure} \begin{centering} \includegraphics[clip,width=0.48\textwidth]{fig6_TcBKT_eb} \par\end{centering} \protect\caption{(color online) The critical BKT transition temperature as a function of the binding energy $E_{b}$ at different in-plane Zeeman fields (a) or out-of-plane Zeeman fields (b). Here and in the next two figures, the color green, red, blue and yellow in the curves denote the superfluid phase gFF, nFF, tnFF and tgFF, respectively.} \label{fig5} \end{figure} We now turn to consider the critical BKT temperature, which is determined by the KT-Nelson criterion, \begin{equation} k_{B}T_{BKT}=\frac{\pi\hbar^{2}}{8m}\left[n_{s,xx}\left(T_{BKT}\right)n_{s,yy}\left(T_{BKT}\right)\right]^{1/2}.\label{eq:BKT2} \end{equation} In the above equation, we have explicitly written down the temperature dependence of the superfluid density, in order to emphasize the fact that the critical BKT temperature should be solved self-consistently. In Figs. \ref{fig5}, \ref{fig6} and \ref{fig7}, we show the results as a function of the binding energy, Zeeman fields and spin-orbit coupling strength, respectively. These results should be regarded as finite-temperature phase diagrams, as they show clearly which kind of superfluid phases is preferable when temperature decreases. In the curves, we use different colors to distinguish different \emph{emerging} superfluid phases: green for the gFF phase, red for the nFF phase, blue for the tnFF phase and finally yellow for the tgFF phase. It is clear that all the four FF superfluid phases have significant critical BKT temperature except for the parameter regime with very small binding energy $E_{b}$ and/or spin-orbit coupling strength $\lambda$, or with very large in-plane Zeeman field $h_{x}$ and/or out-of-plane Zeeman field $h_{z}$. \begin{figure} \begin{centering} \includegraphics[clip,width=0.48\textwidth]{fig5_TcBKT_hzhx} \par\end{centering} \protect\caption{(color online) The critical BKT transition temperature as a function of the out-of-plane Zeeman field $h_{z}$ (a) or the in-plane Zeeman field $h_{x}$ (b).} \label{fig6} \end{figure} As illustrated in Fig. \ref{fig5}, the critical BKT temperature $T_{BKT}$ always increases monotonically with increasing the binding energy $E_{b}$, as the pairing and superfluidity are enhanced at strong interatomic interactions. The binding energy is the dominant factor in forming Cooper pairs. With a small binding energy, the system is mainly of fermionic character. On the contrary, with a sufficiently large binding energy, the system tends to act as a gas of bosons. Therefore, with increasing the binding energy up to some points, the system would lose its fermionic character near the BEC-BCS crossover (i.e., $E_{b}\sim0.5E_{F}$) and hence should become topologically trivial. Indeed, at large binding energy we observe that the system always approaches the topologically trivial gFF phase. The topological phase, either gapless (tnFF in blue) or gapped (tgFF in yellow), is favored at small binding energy, where the critical BKT temperature is lower. Nevertheless, we find that by suitably tuning the parameters, it is possible to have a gapless topological tnFF phase with a sizable critical BKT temperature $T_{BKT}\sim0.09T_{F}$ for the binding energy up to $E_{b}\simeq0.4E_{F}$ (see, for example, the dot-dashed line at the bottom of Fig. \ref{fig5}(a)). This temperature is clearly within the reach in current cold-atom experiments \cite{Ries2014}. On the other hand, the critical BKT temperature decreases monotonically with increasing the Zeeman field, either in-plane or out-of-plane, as shown in Fig. \ref{fig6}. It is readily seen that with decreasing temperature the system would first turn into either the tnFF or tgFF phase at sufficiently large in-plane Zeeman field $h_{x}$ or out-of-plane field $h_{z}$, respectively. While at low Zeeman fields, the topologically trivial gFF phase is preferable. This agrees the observation we made in discussing the low-temperature phase diagrams in Fig. \ref{fig1}. \begin{figure} \begin{centering} \includegraphics[clip,width=0.48\textwidth]{fig7_TcBKT_lambda} \par\end{centering} \protect\caption{(color online) The critical BKT transition temperature as a function of the spin-orbit coupling strength $\lambda$ at different binding energies (a), in-plane Zeeman fields (b) and out-of-plane Zeeman fields (c). } \label{fig7} \end{figure} It is worth noting that, one may use the binding energy dependence or the Zeeman field dependence of the critical BKT temperature to identify different emerging superfluid phases. This is particularly clear for the gapless tnFF and nFF phases, as the curvature of the $T_{BKT}$ curve for those phases behaves quite differently. For the tnFF phase, the curve is concave; while for the nFF phase, it is convex. This change in curvature (i.e, from concave to convex) seems to be related to the local minimum in the superfluid density component $n_{s,xx}$ that we have reported earlier in Fig. \ref{fig3}. We now discuss the critical BKT temperature as a function of the spin-orbit coupling strength $\lambda$, as shown in Fig. \ref{fig7}. Compared with the binding energy dependence and Zeeman field dependence, the dependence of $T_{BKT}$ on the spin-orbit coupling strength is non-monotonic and the emerging superfluid phases can re-appear with increasing the coupling strength. Therefore, the $T_{BKT}$ curve is more subtle to understand. Nevertheless, we may identify that the topologically trivial gFF superfluid phase tends to be favorable at large spin-orbit coupling. This is because the pairing gap is usually enhanced by the spin-orbit coupling, which makes the topological phase transition much more difficult to occur (cf. Eq. (\ref{eq:tpt})). At small spin-orbit coupling, on the other hand, the critical BKT temperature may dramatically decrease to zero, particularly at a small binding energy and/or a large Zeeman field. Thus, for the purpose of observing the gapless topological tnFF phase, experimentally it seems better to use an intermediate spin-orbit coupling strength, i.e., $\lambda\sim E_{F}/k_{F}$. \section{Conclusions} In summary, we have presented a systematic investigation of the Berezinskii-Kosterlitz-Thouless transition in a spin-orbit coupled atomic Fulde-Ferrell superfluid in two dimensions. We have calculated the superfluid density and superfluid transition temperature of various Fulde-Ferrell superfluids. We have paid special attention to an interesting gapless topological Fulde-Ferrell superfluid and have clarified that, by suitably tuning the external parameters - for example, the interatomic interaction strength, in-plane and out-of-plane Zeeman fields, and spin-orbit coupling strength - its observation is within the reach in current cold-atom experimental setups. Our investigation is based on the mean-field theoretical framework, which is supposed to be applicable to a weakly interacting two-dimensional Fermi gas (i.e., the binding energy $E_{b}\leq0.2E_{F}$). For a more reliable and quantitative description, in future studies it would be useful to take into account the strong phase fluctuations by using many-body \textit{T}-matrix theories \cite{Hu2006,Hu2008,Hu2010}. \begin{acknowledgments} X.J.L. and H.H. were supported by the Australian Research Council (ARC) (Grants Nos. FT140100003, FT130100815, DP140103231 and DP140100637) and the National Key Basic Research Special Foundation of China (NKBRSFC-China) (Grant No. 2011CB921502). L.H. was supported by US Department of Energy Nuclear Physics Office. G.L.L. was supported by the National Natural Science Foundation of China (NSFC-China) (Grant Nos. 11175094, 91221205) and the NKBRSFC-China (Grant No. 2011CB921602). \textit{Note added}. Recently, a similar publication by Xu and Zhang became public \cite{Xu2014}. Results qualitatively agree where applicable.\end{acknowledgments}
\section{Introduction} The quest for covering a large number of scenarios beyond the Standard Model (BSM) with current searches for new physics at the LHC requires efficient ways of matching experimental results with theory predictions. A particularly successful approach is the utilization of simplified models \cite{ArkaniHamed:2007fw,Alwall:2008ag,Alves:2011wf,sidewalk}. Recently developed program packages~\cite{Kraml:2013mwa,Papucci:2014rja} provide a convenient framework to employ simplified models for testing BSM theories at the LHC\@. Common to the experimental limits thus obtained is that the simplified models used for data interpretation are characterized by a small number of new particles and a simplified description of particle production and decay. The underlying assumption for this treatment is that the more model-specific details of the production and decay dynamics have little influence on the signal efficiencies. In this study we question the validity of this assumption. We focus on light flavor squark production in the minimal supersymmetric model with $R$-parity conservation. In supersymmetric models, squark production proceeds also through the exchange of a gluino in the $t$-channel and thus includes various processes with left- and right-chiral squarks and anti-squarks in the final state, $pp\to \s{q}_i^{}\s{q}_j^*$ and $pp\to \s{q}_i\s{q}_j$, with the chirality $i,j = \text{L},\text{R}$. In the simplified models adopted by the ATLAS and CMS collaborations, however, the usual choice is to decouple the gluino, see, e.g., \cite{Aad:2014wea, Chatrchyan:2013lya,Chatrchyan:2014lfa}. Consequently, the only contributing production mode is $pp\to{\widetilde{q}}_i^{}{\widetilde{q}}_i^*$. As $pp\to{\widetilde{q}}_i^{}{\widetilde{q}}_i^*$ is in general not the dominant production channel, it is an important question whether the signal efficiencies -- and hence the resulting exclusion limits -- derived from this production mode are applicable to the more general case of supersymmetric models. In this paper we compare the efficiencies for squark production and the exclusion limits for squark masses obtained by using a simplified model with those obtained in the complete supersymmetric model. We focus on two representative all-hadronic analyses performed by CMS: one based on the discriminating variable $\alpha_{\text{T}}$~\cite{Chatrchyan:2013lya}, and one based on missing transverse energy, denoted by $\slashed{H}_{\text{T}}$~\cite{Chatrchyan:2014lfa}. As in the experimental analyses, we assume a direct decay of the squark into the neutralino. The respective topology is often referred to as T2~\cite{Chatrchyan:2013sza}. The simplified model with a decoupled gluino is denoted by $\text{T2}_\infty$ in the following, while the complete supersymmetric model with a finite gluino mass is denoted by $\text{T2}_\mgo$. The simplified model commonly adopted by ATLAS and CMS for the hadronic searches for squarks corresponds to $\text{T2}_\infty$. We find very large deviations between the efficiencies for $\text{T2}_\infty$ and $\text{T2}_\mgo$ in certain regions of parameter space, in particular for large squark masses. However, we show that for the parameter region relevant for the current exclusion limits, the differences turn out to be moderate, and the exclusion limits obtained from applying the efficiencies for $\text{T2}_\infty$ are close to the ones for $\text{T2}_\mgo$. Comparing the differences to the theoretical uncertainties of the cross section normalization, we find that both are of the same size for large regions in the considered parameter space. In addition, we compare the two analyses based on $\alpha_{\text{T}}$ and $\slashed{H}_{\text{T}}$ and show that the $\alpha_{\text{T}}$ variable is less sensitive to the difference in the production dynamics and modes. Furthermore, whereas in the $\slashed{H}_{\text{T}}$ analysis the $\text{T2}_\infty$ efficiencies lead to an overestimation of the exclusion limits, in the $\alpha_{\text{T}}$ analysis $\text{T2}_\infty$ yields conservative limits. This paper is structured as follows. In section \ref{sec:production} we review the different squark production processes and their dependence on the gluino mass. In section \ref{sec:analysis} we introduce the two CMS searches and the parameter scan. The results for the comparison of efficiencies and squark mass limits are presented in section \ref{sec:results}. We conclude in section~\ref{sec:conclusion}. \section{Production processes of squarks at the LHC} \label{sec:production} The production of strongly interacting sparticles is an important production channel for probing supersymmetry at the LHC. The strength and the kinematic distribution of squark (anti-)squark production not only depend on the mass spectrum of the squarks, but also on the gluino mass ${m_{\widetilde{g}}}$, since all squark production processes -- except for the case $pp\to \s{q}_i^{}\s{q}_i^*$, $i=\text{L},\text{R}$ -- require a $t$-channel gluino exchange (cf. the diagrams in figure \ref{fig:sqsqbar_production}). \begin{figure}[htp] \centering \setlength{\unitlength}{1\textwidth} \begin{picture}(0.9,0.34) \put(-0.024,0.18){ \put(-0.024,0.06){ a)} ~~~\includegraphics[width=3.2cm]{pdffig/nsusy1.pdf}~~ \includegraphics[width=3.2cm]{pdffig/nsusy2.pdf}~~ \includegraphics[width=3.2cm]{pdffig/nsusy3.pdf}~~ \includegraphics[width=3.2cm]{pdffig/nsusy4.pdf}~~ } \put(-0.024,0.0) \put(-0.024,0.06){ b)} ~~~\includegraphics[width=3.2cm]{pdffig/nsusy5.pdf}~~~~~~ \includegraphics[width=3.2cm]{pdffig/nsusy6.pdf} } \end{picture} \caption{a)~Production process $pp\rightarrow \widetilde{q}_i^{}\widetilde{q}^*_i$ with $i=\text{L},\text{R}$ with a decoupled gluino. For $i=\text{L}$, these graphs contribute to $\text{T2}_\infty$. b)~Diagrams for $pp\rightarrow \widetilde{q}_i^{}\widetilde{q}^*_j$ and $pp\rightarrow \widetilde{q}_i^{}\widetilde{q}_j^{}$, with $i,j=\text{L},\text{R}$. These are present if the gluino is not decoupled. Thus both types of diagrams a) and b) contribute to $\text{T2}_\mgo$.} \label{fig:sqsqbar_production} \end{figure} We first discuss the relative importance of different squark production processes as a function of the mass ratio ${m_{\widetilde{g}}}/{m_{\widetilde{q}}}$, where we make the convenient assumption of a common mass ${m_{\widetilde{q}}}$ for the squarks of the first and second generation.\footnote{% Here we consider ${m_{\widetilde{g}}}$ and ${m_{\widetilde{q}}}$ as free parameters defined at the TeV scale. Note, however, that a large ratio ${m_{\widetilde{g}}}/{m_{\widetilde{q}}}$ is inaccessible in models where the supersymmetry breaking is mediated at a high scale. For a detailed discussion see, e.g., \cite{Jaeckel:2011wp}.} The third generation is not considered here. While varying the ratio ${m_{\widetilde{g}}}/{m_{\widetilde{q}}}$, we kept the total production cross section of squarks and gluinos, denoted by $\sigma_{\{{\widetilde{g}}{\widetilde{q}}\}}$, fixed. This requirement determines ${m_{\widetilde{q}}}$ and ${m_{\widetilde{g}}}$. We show relative cross section contributions for $\sigma_{\{{\widetilde{g}}{\widetilde{q}}\}}\simeq 1000\ensuremath{\,\textnormal{fb}}$ and $\sigma_{\{{\widetilde{g}}{\widetilde{q}}\}}\simeq 10\ensuremath{\,\textnormal{fb}}$. These values represent typical cross section upper limits from the 8\,TeV LHC null-search for regions with very small and very high sensitivities, respectively. For ${m_{\widetilde{g}}}/{m_{\widetilde{q}}}\gtrsim3$, this corresponds to squark masses of about $500\text{\,GeV}$ and $1\text{\,TeV}$, respectively. We computed the total production cross section at NLO with \textsc{Prospino} \cite{Beenakker:1996ch}, while the individual contributions were calculated at LO with \textsc{MadGraph}~5~\cite{Alwall:2011uj}\@. \begin{figure}[t] \centering \setlength{\unitlength}{1\textwidth} \begin{picture}(0.893,0.354) \put(-0.024,0.0){ \put(0.0,0.025){\includegraphics[width=0.45\textwidth]{pdffig/contributions_main_1000fb_wq2g.pdf}} \put(0.2,0.0){\footnotesize ${m_{\widetilde{g}}}/{m_{\widetilde{q}}}$} \put(-0.03,0.14){\rotatebox{90}{\footnotesize $\sigma_i/\sigma_{\{{\widetilde{g}}{\widetilde{q}}\}}$}} \put(0.26,0.256){\tiny ${\widetilde{q}}_1^{(\!*\!)}\!{\widetilde{q}}_1^{(\!*\!)}$} \put(0.177,0.22){\tiny ${\widetilde{g}}{\widetilde{q}}_1^{(\!*\!)}$} \put(0.115,0.184){\tiny ${\widetilde{g}}\go$} \put(0.36,0.142){\tiny ${\widetilde{q}}_1^{(\!*\!)}\!{\widetilde{q}}_2^{(\!*\!)}$} \put(0.334,0.224){\tiny ${\widetilde{q}}_2^{(\!*\!)}\!{\widetilde{q}}_2^{(\!*\!)}$} \put(0.048,0.078){\tiny ${\widetilde{g}}{\widetilde{q}}_2^{(\!*\!)}$} % \put(0.04,0.313){\footnotesize $\sigma_{\{{\widetilde{g}}{\widetilde{q}}\}}\simeq1000\ensuremath{\,\textnormal{fb}}$} } \put(0.492,0.0){ \put(0.0,0.025){\includegraphics[width=0.45\textwidth]{pdffig/contributions_main_10fb.pdf}} \put(0.2,0.0){\footnotesize ${m_{\widetilde{g}}}/{m_{\widetilde{q}}}$} \put(-0.03,0.14){\rotatebox{90}{\footnotesize $\sigma_i/\sigma_{\{{\widetilde{g}}{\widetilde{q}}\}}$}} \put(0.26,0.263){\tiny ${\widetilde{q}}_1^{(\!*\!)}{\widetilde{q}}_1^{(\!*\!)}$} \put(0.155,0.193){\tiny ${\widetilde{g}}{\widetilde{q}}_1^{(\!*\!)}$} \put(0.112,0.158){\tiny ${\widetilde{g}}\go$} \put(0.24,0.155){\tiny ${\widetilde{q}}_1^{(\!*\!)}{\widetilde{q}}_2^{(\!*\!)}$} \put(0.357,0.191){\tiny ${\widetilde{q}}_2^{(\!*\!)}{\widetilde{q}}_2^{(\!*\!)}$} % \put(0.04,0.313){\footnotesize $\sigma_{\{{\widetilde{g}}{\widetilde{q}}\}}\simeq10\ensuremath{\,\textnormal{fb}}$} } \end{picture} \caption{ Relative contributions to the production cross section of squarks and gluinos as a function of the ratio ${m_{\widetilde{g}}}/{m_{\widetilde{q}}}$ along iso-cross section curves for the $8\text{\,TeV}$ LHC\@. \emph{Left panel:}~For a total production cross section $\sigma_{\{{\widetilde{g}}{\widetilde{q}}\}}\simeq 1000\ensuremath{\,\textnormal{fb}}$, i.e., small ${m_{\widetilde{g}}}$ and ${m_{\widetilde{q}}}$. \emph{Right panel:}~For $\sigma_{\{{\widetilde{g}}{\widetilde{q}}\}}\simeq 10\ensuremath{\,\textnormal{fb}}$, i.e., large ${m_{\widetilde{g}}}$ and ${m_{\widetilde{q}}}$. We take into account all production mechanisms of gluinos and of squarks of the first and second generation. These squarks are assumed to have a common mass ${m_{\widetilde{q}}}$. Here we denote ${\widetilde{q}}^{(*)}_1$, ${\widetilde{q}}^{(*)}_2$ to be all first and second generation (anti)squarks not distinguishing between left- and right superpartners. } \label{fig:contri1} \end{figure} \begin{figure}[h!] \centering \setlength{\unitlength}{1\textwidth} \begin{picture}(0.893,0.35) \put(-0.024,0){ \put(0.0,0.025){\includegraphics[width=0.45\textwidth]{pdffig/contributions_sub_1stgen_xs1000.pdf}} \put(0.2,0.0){\footnotesize ${m_{\widetilde{g}}}/{m_{\widetilde{q}}}$} \put(-0.03,0.14){\rotatebox{90}{\footnotesize $\sigma_i/\sigma_{{\widetilde{q}}_1{\widetilde{q}}_1}$}} \put(0.2,0.254){\tiny ${\widetilde{q}}_\L{\widetilde{q}}_\L\!+\!{\widetilde{q}}_{\text{R}}{\widetilde{q}}_{\text{R}}$} \put(0.12,0.227){\tiny ${\widetilde{q}}_\L{\widetilde{q}}_{\text{R}}$} \put(0.1,0.115){\tiny ${\widetilde{q}}_\L^{}{\widetilde{q}}_\L^*\!+\!{\widetilde{q}}_{\text{R}}^{}{\widetilde{q}}_{\text{R}}^*$} \put(0.33,0.163){\tiny ${\widetilde{q}}_\L^{}{\widetilde{q}}_{\text{R}}^* \!+\! {\widetilde{q}}_{\text{R}}^{}{\widetilde{q}}_\L^*$} % \put(0.01,0.313){\footnotesize Subcontributions of ${\widetilde{q}}_1^{(*)}{\widetilde{q}}_1^{(*)}$, $\sigma_{\{{\widetilde{g}}{\widetilde{q}}\}}\simeq1000\ensuremath{\,\textnormal{fb}}$} } \put(0.492,0){ \put(0.0,0.025){\includegraphics[width=0.45\textwidth]{pdffig/contributions_sub_1stgen_xs10.pdf}} \put(0.2,0.0){\footnotesize ${m_{\widetilde{g}}}/{m_{\widetilde{q}}}$} \put(-0.03,0.14){\rotatebox{90}{\footnotesize $\sigma_i/\sigma_{{\widetilde{q}}_1{\widetilde{q}}_1}$}} \put(0.2,0.265){\tiny ${\widetilde{q}}_\L{\widetilde{q}}_\L\!+\!{\widetilde{q}}_{\text{R}}{\widetilde{q}}_{\text{R}}$} \put(0.1,0.225){\tiny ${\widetilde{q}}_\L{\widetilde{q}}_{\text{R}}$} \put(0.105,0.086){\tiny ${\widetilde{q}}_\L^{}{\widetilde{q}}_\L^*\!+\!{\widetilde{q}}_{\text{R}}^{}{\widetilde{q}}_{\text{R}}^*$} \put(0.05,0.148){\tiny ${\widetilde{q}}_\L^{}{\widetilde{q}}_{\text{R}}^* \!+\! {\widetilde{q}}_{\text{R}}^{}{\widetilde{q}}_\L^*$} % \put(0.029,0.313){\footnotesize Subcontributions of ${\widetilde{q}}_1^{(*)}{\widetilde{q}}_1^{(*)}$, $\sigma_{\{{\widetilde{g}}{\widetilde{q}}\}}\simeq10\ensuremath{\,\textnormal{fb}}$} } \end{picture} \caption{ Relative contributions to the production cross section of squarks of the first generation at the $8\text{\,TeV}$ LHC\@. These are the subcontributions of ${\widetilde{q}}_1^{(*)}{\widetilde{q}}_1^{(*)}$ in figure~\protect\ref{fig:contri1} as a function of the ratio ${m_{\widetilde{g}}}/{m_{\widetilde{q}}}$ along iso-cross section curves. \emph{Left panel:}~For a total production cross section $\sigma_{\{{\widetilde{g}}{\widetilde{q}}\}}\simeq 1\ensuremath{\,\textnormal{pb}}$, corresponding to small squark masses. \emph{Right panel:}~For $\sigma_{\{{\widetilde{g}}{\widetilde{q}}\}}\simeq 10\ensuremath{\,\textnormal{fb}}$, corresponding to large squark masses.} \label{fig:contri2} \end{figure} In figure \ref{fig:contri1} we show the relative strength of all possible combinations of first generation squark production, ${\widetilde{q}}_1^{(*)}$, second generation squark production, ${\widetilde{q}}_2^{(*)}$, and gluino production. In figure~\ref{fig:contri2}, the subcontributions to first generation squark pair production, ${\widetilde{q}}_1^{(*)}{\widetilde{q}}_1^{(*)}$, are shown. We summed processes that give equal contributions, such as ${\widetilde{q}}_\L^{}{\widetilde{q}}_\L^{}$ and ${\widetilde{q}}_{\text{R}}^{}{\widetilde{q}}_{\text{R}}^{}$. Subcontributions that are not displayed (antisquark pair production, for instance) are below the range displayed here. From figures \ref{fig:contri1} and \ref{fig:contri2} we can draw several conclusions. First, the production of first generation squark pairs, ${\widetilde{q}}_1^{(*)}{\widetilde{q}}_1^{(*)}$, is the dominant production channel over a large range of ${m_{\widetilde{g}}}/{m_{\widetilde{q}}}$. This dominance is even more pronounced for $\sigma_{\{{\widetilde{g}}{\widetilde{q}}\}}\simeq 10\ensuremath{\,\textnormal{fb}}$, i.e., for larger ${m_{\widetilde{q}}}$. Second, among its subcontributions, the channel ${\widetilde{q}}_\L^{}{\widetilde{q}}_\L^{}+{\widetilde{q}}_{\text{R}}^{}{\widetilde{q}}_{\text{R}}^{}$ (with ${\widetilde{q}} \equiv {\widetilde{q}}_1=\s u, \s d\,$) is the dominant channel. Although for large ${m_{\widetilde{g}}}/{m_{\widetilde{q}}}$ its contribution is suppressed through the gluino mass appearing in the $t$-channel propagator, the relative contributions stays dominant up to ${m_{\widetilde{g}}}\simeq7{m_{\widetilde{q}}}$ (for $\sigma_{\{{\widetilde{g}}{\widetilde{q}}\}}\simeq 1000\ensuremath{\,\textnormal{fb}}$) and above ${m_{\widetilde{g}}}/{m_{\widetilde{q}}}=10$ (for $\sigma_{\{{\widetilde{g}}{\widetilde{q}}\}}\simeq 10\ensuremath{\,\textnormal{fb}}$). These effects are due to the relatively large parton luminosities for $uu$ and $dd$ initial states when approaching large partonic center-of-mass energies in the hard scattering process. In contrast, although equally enhanced through a higher parton luminosity, the relative contribution of ${\widetilde{q}}_\L^{}{\widetilde{q}}_{\text{R}}^{}$ decreases rapidly with increasing ${m_{\widetilde{g}}}/{m_{\widetilde{q}}}$. This is because its cross section is suppressed by $1/{m_{\widetilde{g}}}^2$ compared to ${\widetilde{q}}_\L^{}{\widetilde{q}}_\L^{}$ and ${\widetilde{q}}_{\text{R}}^{}{\widetilde{q}}_{\text{R}}^{}$ in the limit of large ${m_{\widetilde{g}}}$. The subcontributions to second generation squark pair production are completely dominated by ${\widetilde{q}}_\L^{}{\widetilde{q}}^*_\L$ and ${\widetilde{q}}_{\text{R}}^{}{\widetilde{q}}^*_{\text{R}}$. Its absolute contribution is dominated by the diagrams in the first row of figure~\ref{fig:sqsqbar_production}, which are independent of the gluino mass as well as independent of the squark flavor. In summary, even for ${m_{\widetilde{g}}}/{m_{\widetilde{q}}}$ as large as 10 the channels ${\widetilde{q}}_\L^{}{\widetilde{q}}^*_\L$, ${\widetilde{q}}_{\text{R}}^{}{\widetilde{q}}^*_{\text{R}}$ are not necessarily the dominant production mode of squarks, and it is crucial to take into account other channels, in particular ${\widetilde{q}}_\L^{}{\widetilde{q}}_\L^{}$ and ${\widetilde{q}}_{\text{R}}^{}{\widetilde{q}}_{\text{R}}^{}$. These, in turn, yield different event kinematics and thus, in principle, different signal efficiencies for the experimental searches. \section{Analyses and parameter scan} \label{sec:analysis} We consider two representative all-hadronic analyses performed by the CMS collaboration using the $8\text{\,TeV}$ LHC data set. One analysis is based on the discriminating variable $\alpha_{\text{T}}$~\cite{Chatrchyan:2013lya}, and one is based on missing transverse energy $\slashed{H}_{\text{T}}$ \cite{Chatrchyan:2014lfa}. Among other models, these searches were interpreted in terms of the simplified model $\text{T2}_\infty$, which consists of squark production via the process $pp\to{\widetilde{q}}\sq^*\;({m_{\widetilde{g}}}\to\infty)$, followed by a squark decay into a neutralino $\s{\chi}^0_1$ (the lightest supersymmetric particle, LSP, in this model) and a quark with branching ratio 1. \subsection{All-hadronic analyses} \label{sec:analyses} The CMS analysis \cite{Chatrchyan:2014lfa} is sensitive to the pair production of squarks (and gluinos) decaying into one (two) jets and an LSP. The search requires at least three jets, and the data is divided into several categories (3--5, 6--7, 8 jets). Consequently, the analysis is more sensitive to final states resulting from longer cascade decays of gluinos and squarks. Since we are interested in the pair production of squarks, we concentrate on the 3--5 jet analysis. The search is based on the two variables \begin{align} H_{\text{T}}&= \sum_{j\,\in\,\text{jets}} p^j_{\text{T}}&& \text{for jets with} \quad p^j_{\text{T}}>50 \text{\,GeV}\,,\; |\eta_j|<2.5\,, \\ \slashed{H}_{\text{T}}& =\left|\slashed{\vec{H}}_{\text{T}}\right|=\left|-\sum_{j\,\in\,\text{jets}}\vec{p}^j_{\text{T}}\right| &&\text{for jets with} \quad p^j_{\text{T}}>30 \text{\,GeV}\,,\; |\eta_j|<5\,. \label{eq:htmissdef} \end{align} $H_{\text{T}}$ characterizes the total visible hadronic activity and $\slashed{H}_{\text{T}}$ the momentum imbalance in an event. The selection regions are categorized as summarized in table \ref{tab:MHT}. \renewcommand{\arraystretch}{1.2} \begin{table}[bt] \centering \begin{tabular}{crl} \toprule \multicolumn{1}{l}{bins} & \multicolumn{1}{l}{$H_{\text{T}}\; [\text{\,GeV}\,]$} & $ \slashed{H}_{\text{T}} \;[\text{\,GeV}\,]$\\ \midrule 1\,--\,4\phantom{1} & \phantom{1}500\,--\,800\phantom{1} & 200\,--\,300;~300\,--\,450;~450\,--\,600;~\textgreater\,600 \\ 5\,--\,8\phantom{1} & \phantom{1}800\,--\,1000 & 200\,--\,300;~300\,--\,450;~450\,--\,600;~\textgreater\,600 \\ 9\,--\,12 & 1000\,--\,1250 & 200\,--\,300;~300\,--\,450;~450\,--\,600;~\textgreater\,600 \\ 13\,--\,15 & 1250\,--\,1500 & 200\,--\,300;~300\,--\,450;~\textgreater 450 \\ 16\,--\,17 & \textgreater 1500 & 200\,--\,300;~\textgreater\,300 \\ \multicolumn{1}{c}{0} & \textgreater \phantom{1}500 &\textgreater\,200 \\ \bottomrule \end{tabular} \caption{Categorization of selection regions (bins) for the $\slashed{H}_{\text{T}}$ analysis \cite{Chatrchyan:2014lfa}.} \label{tab:MHT} \end{table} Further cuts are $\left|\Delta\phi(\vec{p}_{j},\slashed{\vec{H}}_{\text{T}})\right|>0.5$ for the two hardest jets and $\left|\Delta\phi(\vec{p}_{j},\slashed{\vec{H}}_{\text{T}})\right|>0.3$ for the third hardest jet, as well as a veto against isolated electrons and muons with $p_{\text{T}}>10\text{\,GeV}$. For details see \cite{Chatrchyan:2014lfa}. We refer to this analysis in the following as the $\slashed{H}_{\text{T}}$ (MHT) analysis.\\ Another CMS analysis \cite{Chatrchyan:2013lya} is based on the variable $\alpha_{\text{T}}$, which is a powerful variable for discrimination against QCD multijet background. It rejects multijet events without significant $\slashed{E}_{\text{T}}$ \cite{Randall:2008rw,Khachatryan:2011tk}. For a dijet system, $\alpha_{\text{T}}$ is defined as \begin{align} &\alpha_{\text{T}}=\frac{E_{\text{T}}^{j_2}}{M_{\text{T}}}\,, && M_{\text{T}}=\sqrt{\left(\sum_{i=1}^2E_{\text{T}}^{j_i}\right)^2-\left(\sum_{i=1}^2p_x^{j_i}\right)^2-\left(\sum_{i=1}^2p_y^{j_i}\right)^2}, \end{align} where $j_2$ denotes the less energetic jet. $\alpha_{\text{T}}$ is 0.5 for perfectly measured dijet events that are back-to-back in $\phi$. If the two jets are not back-to-back and recoil against a large $\slashed{E}_{\text{T}}$, $\alpha_{\text{T}}$ becomes larger than $0.5$. For suppression of mismeasured QCD background, $\alpha_{\text{T}}$ is required to be larger than 0.55. For events with more than two jets, a pseudo-dijet system is formed, such that the absolute $E_{\text{T}}$ difference, denoted as $\Delta H_{\text{T}}$, between the two pseudo-jets is minimized. In this case, the variable is generalized to \begin{align} &\alpha_{\text{T}}= \frac{1}{2}\frac{H_{\text{T}}-\Delta H_{\text{T}}}{\sqrt{H_{\text{T}}^2-\slashed{H}_{\text{T}}^2}}, \end{align} where $H_{\text{T}}$ is the scalar sum of the transverse energies $E_{\text{T}}$ of the jets, $H_{\text{T}}=\sum_j E_{\text{T}}^{j}$,\footnote{Note that this definition differs from the one used in the MHT analysis.} and $\slashed{H}_{\text{T}}$ is as defined in (\ref{eq:htmissdef}). As we consider light flavor squark-squark production, we choose the signal region using 2--3 jets without a $b$-tagged jet. To maximize the sensitivity, the events are divided in eight bins based on $H_{\text{T}}$: one bin in the range of 275--325\text{\,GeV}, one in the range of 325--375\text{\,GeV}, five bins between 375--875\text{\,GeV}\ in steps of 100\text{\,GeV}, and an open bin $>875$\text{\,GeV}. In the following we will refer to these eight bins as bin\,$1\dots 8$\, and to the combination of all bins as bin\,0. The selection criteria for jets in the first two bins differ from the others: for bin\,1 (bin\,2), $E_{\text{T}}^j >37\text{\,GeV}$ $(43\text{\,GeV})$ is required. In these same two bins, the two highest-$E_{\text{T}}$ jets are required to have $E_{\text{T}}^j >73\text{\,GeV}$ $(87\text{\,GeV})$. For the other bins, the threshold for the two highest-$E_{\text{T}}$ jets is $E_{\text{T}}^j>100\text{\,GeV}$ each, and all jets are required to have $E_{\text{T}}^j>50\text{\,GeV}$. In addition, for all bins jets are required to have $|\eta|<3$ ($|\eta|<2.5$ for the highest-$E_{\text{T}}$ jet). For details see \cite{Chatrchyan:2013lya}. We refer to this analysis in the following as the $\alpha_{\text{T}}$ analysis. \subsection{Event generation and parameter scan} In order to compute signal efficiencies, we performed a Monte Carlo event simulation. We used \textsc{MadGraph}~5~\cite{Alwall:2011uj} to simulate the hard scattering of the squarks, whereafter \textsc{Pythia}~6~\cite{Sjostrand:2006za} was used for the decays of the squarks into neutralinos as well as for showering and hadronization. As the MHT analysis requires at least three hard jets, in our model at least one jet has to arise from initial state radiation. We therefore include up to one jet in the hard scattering and performed an MLM matching \cite{alwall-2008-53} with initial state radiation from \textsc{Pythia}~6. We chose a matching scale $Q_{\text{cut}}=46\text{\,GeV}$ and $p_{\text{T}}^{\text{jet}}>30\text{\,GeV}$ in \textsc{MadGraph}~5. The gluino and squark\footnote{% We assumed a pure bino for the computation of squark widths.} widths were computed with \textsc{MadGraph}~5. The branching ratios of squark decays to the neutralino were set to 1. We used \textsc{Delphes}~3.0.11~\cite{deFavereau:2013fsa} with \textsc{Fastjet}~\cite{fastjet} for detector simulation using the standard CMS settings included in this version of \textsc{Delphes}, but changed the $b$-tag misidentification rate to 0.01 \cite{Chatrchyan:2013lya,Chatrchyan:2012jua}. We performed parameter scans in the mass region $m_{\s\chi^0_1}=100\dots1400\text{\,GeV}$ and ${m_{\widetilde{q}}}=500\dots1500 \text{\,GeV}$ in steps of $100\text{\,GeV}$, with ${m_{\widetilde{q}}}-m_{\s\chi^0_1}\geq 100\text{\,GeV}$, for two mass ratios ${m_{\widetilde{g}}}/{m_{\widetilde{q}}}=2,4$. For both ratios, associated gluino and gluino pair production are negligible. We computed the efficiencies for MHT and $\alpha_{\text{T}}$ for all sub-channels of first generation squark production (as listed in section \ref{sec:production}), as well as for ${\widetilde{q}}_1^{(*)}{\widetilde{q}}_2^{(*)}+{\widetilde{q}}_2^{(*)}{\widetilde{q}}_2^{(*)}$, or production of at least one second generation squark (or anti-squark). Finally, as a reference we computed the efficiencies for $pp\to {\widetilde{q}}_\L^{}{\widetilde{q}}_\L^*$ for the case ${m_{\widetilde{g}}}\to\infty$, which amounts to the simplified model $\text{T2}_\infty$. \section{Results} \label{sec:results} In this section, we show the deviations from the simplified model $\text{T2}_\infty$ that arise from different production mechanisms, both at the efficiency level (see section \ref{section:efficiencies}) and at the level of squark and LSP mass limits (see section \ref{sec:limits}). We find that although there are some substantial differences at the efficiency level, the deviations in the mass limits based on current LHC data are rather moderate. \subsection{Efficiencies} \label{section:efficiencies} Comparing the signal acceptance\,$\times$\,efficiency $A\epsilon$ (simply called ``efficiency'' in the following) for $\text{T2}_\infty$ and the other individual production mechanisms $M$ that contribute to $\text{T2}_\mgo$, for individual bins\,$i$ we found very large relative deviations $$\frac{A\epsilon^i(M)-A\epsilon^i(\text{T2}_\infty)}{A\epsilon^i(\text{T2}_\infty)},$$ that were up to about 220\%. However, not all bins are equally relevant for setting exclusion limits. In order to take this into account, we consider here the most sensitive bin only, which we define as the bin that yields the largest ratio $S^i/S^i_{95\%\,\text{C.L.}}(B)$, where $S^i\propto A\epsilon^i$ is the expected number of signal events in bin\,$i$, while $S^i_{95\%\,\text{C.L.}}(B)$ is the corresponding required number of signal events that would provide a 95\%\,C.L. exclusion if the data would equal the background prediction $B$. Of particular interest is the production mechanism $pp\rightarrow \widetilde{q}_\L^{}\widetilde{q}_\L^{},{\widetilde{q}}_{\text{R}}^{}{\widetilde{q}}_{\text{R}}^{}$, as it is the dominant contribution (see section \ref{sec:production}). The deviations in the efficiencies obtained for the production mode $pp\to{\widetilde{q}}_\L^{}{\widetilde{q}}_\L^{}$ from those obtained for $\text{T2}_\infty$ are shown in figure \ref{fig:accimportant} for ${m_{\widetilde{g}}}/{m_{\widetilde{q}}}=2$. \subsubsection*{MHT analysis} An overall feature for the MHT analysis is that the $A \epsilon$ for $M=\widetilde{q}_\L^{}\widetilde{q}_\L^{}$ are smaller than those for $\text{T2}_\infty$. The deviations range from up to $- 70\%$ in the region where $m_{\widetilde{q}}-m_{\s\chi^0_1}\approx 100\text{\,GeV}$, to a few percent for small LSP and squark masses. For most masses bin\,12 or bin\,17 (for large squark and small LSP masses) are the most sensitive bins. Another important production mode is $\widetilde{q}_\L^{}\widetilde{q}_{\text{R}}^*$ which contributes significantly to the total cross section for $m_{\widetilde{g}}/m_{\widetilde{q}}\approx 2-5$. For $m_{\widetilde{g}}/m_{\widetilde{q}}=2$, the deviations are largest for $m_{\widetilde{q}}-m_{\s\chi^0_1}\approx 100$\,GeV and reach up to $-80$\%. The deviations decrease to $-5$\% in the region $m_{\widetilde{q}}-m_{\s\chi^0_1}\approx 700$\,GeV and increase again up to $-20$\% for large squark and small LSP masses. For gluino masses $m_{\widetilde{g}}/m_{\widetilde{q}}=4$, the general features remain the same, but the absolute values of the deviations increase by a few percentage points. All deviations are negative, meaning that the $A\epsilon$ for $\text{T2}_\infty$ is larger then for $\widetilde{q}_\L^{}\widetilde{q}_{\text{R}}^*$. For $\widetilde{q}_\L^{}\widetilde{q}_{\text{R}}^{}$ we found deviations as large as 220\%. However, the contribution of $\widetilde{q}_\L^{}\widetilde{q}_{\text{R}}^{}$ to the total cross section is completely negligible for ${m_{\widetilde{g}}}/{m_{\widetilde{q}}}=4$ and rather small for ${m_{\widetilde{g}}}/{m_{\widetilde{q}}}=2$. \begin{figure}[h!] \centering \setlength{\unitlength}{1\textwidth} \begin{picture}(0.85,1.32) \put(0.,-0.006){ \put(0.0,0.025){\includegraphics[scale=0.61]{pdffig/qLqL2_aT_diff.pdf}} \put(0.12,0.56){${\widetilde{q}}_\L{\widetilde{q}}_\L$, ${m_{\widetilde{g}}}/{m_{\widetilde{q}}}=2$, $\alpha_{\text{T}}$} \put(0.38,0.02){\footnotesize ${m_{\widetilde{q}}}\,[\text{\,GeV}\,]$} \put(-0.013,0.286){\rotatebox{90}{\footnotesize $m_{\s\chi^0_1}\,[\text{\,GeV}\,]$}} \put(0.835,0.525){\rotatebox{-90}{\footnotesize $\left(A\epsilon({\widetilde{q}}_\L{\widetilde{q}}_\L)-A\epsilon(\text{T2}_\infty)\right)/A\epsilon(\text{T2}_\infty)\;[\,\%\,]$}} } \put(0,0.64){ \put(0.0,0.025){\includegraphics[scale=0.61]{pdffig/qLqL2_MHT_diff.pdf}} \put(0.12,0.56){${\widetilde{q}}_\L{\widetilde{q}}_\L$, ${m_{\widetilde{g}}}/{m_{\widetilde{q}}}=2$, MHT} \put(0.38,0.02){\footnotesize ${m_{\widetilde{q}}}\,[\text{\,GeV}\,]$} \put(-0.013,0.286){\rotatebox{90}{\footnotesize $m_{\s\chi^0_1}\,[\text{\,GeV}\,]$}} \put(0.835,0.525){\rotatebox{-90}{\footnotesize $\left(A\epsilon({\widetilde{q}}_\L{\widetilde{q}}_\L)-A\epsilon(\text{T2}_\infty)\right)/A\epsilon(\text{T2}_\infty)\;[\,\%\,]$}} } \end{picture} \caption{Relative difference of signal efficiencies $A\epsilon$ for ${\widetilde{q}}_\L^{}{\widetilde{q}}_\L^{}$-production with ${m_{\widetilde{g}}}/{m_{\widetilde{q}}}=2$ and ${\widetilde{q}}_\L^{}{\widetilde{q}}_\L^{\,*}$-production with ${m_{\widetilde{g}}}/{m_{\widetilde{q}}}\to\infty$ ($\text{T2}_\infty$) for the MHT analysis (upper panel) and the $\alpha_{\text{T}}$-analysis (lower panel) at the $8\text{\,TeV}$ LHC\@. The relative difference is denoted by the color code and given in percent. The error corresponds to the Monte Carlo error from event generation. For each mass point we compare the efficiencies of the bin that yields the largest sensitivity within the $\text{T2}_\infty$ model. The corresponding bin is displayed below the relative difference. } \label{fig:accimportant} \end{figure} \clearpage \subsubsection*{$\alpha_{\text{T}}$ analysis} The $\alpha_{\text{T}}$ analysis turns out to be more robust towards the different production and gluino mass scenarios. The efficiencies for $\widetilde{q}_\L^{}\widetilde{q}_\L^{}$ are mainly larger than for $\text{T2}_\infty$ (with the exception of small squark/LSP masses and mass differences). The deviations are in the range of $+20$ to $-40$\%. The most sensitive bin for large squark and small neutralino masses is bin\,8. The deviations of the production mode $\widetilde{q}_\L^{}\widetilde{q}_\L^*$ from $\text{T2}_\infty$ are around a few percent. The largest deviations occur for small squark and LSP masses with around $-10$\%. For larger $m_{\widetilde{g}}/m_{\widetilde{q}}=4$, the deviations become even smaller, as expected. The $A\epsilon$ for $\widetilde{q}_\L^{}\widetilde{q}_{\text{R}}^*$ are larger than for $\text{T2}_\infty$. The largest deviations are $\approx 30$\% for large squark masses. \subsection{Exclusion limits} \label{sec:limits} Our final goal is to examine the effect of the different production mechanisms and gluino masses on the search sensitivity. From the efficiencies we derived 95\% C.L.\ exclusion limits in the ${m_{\widetilde{q}}}$-$m_{\s\chi^0_1}$ plane for ${m_{\widetilde{g}}}/{m_{\widetilde{q}}}=2,4$, for both the simplified model $\text{T2}_\infty$ and the complete supersymmetric model $\text{T2}_\mgo$. In order to obtain the efficiencies for $\text{T2}_\mgo$ from the individual production channels, we compute the weighted mean of the efficiencies, where we use NLO cross sections computed with \textsc{Prospino\,2}~\cite{Beenakker:1996ch} for the weights. For the determination of the most sensitive bin for $\text{T2}_\mgo$ and $\text{T2}_\infty$, we included bin\,0 (the sum of all bins of the analysis). The most sensitive bin is determined, as before, on the basis of background expectation only. To obtain the exclusion limits, we compute \begin{align} \mu^{-1}=\left(\frac{(A\epsilon)^i\times \int\! \mathcal{L} \times\sigma_{\text{tot}} }{S^i_{95\%\text{CL}}(\text{data})}\right)_{i\,=\,\text{most sensitive bin}}, \end{align} for each point in the ${m_{\widetilde{q}}}$-$m_{\s\chi^0_1}$ plane. Here, $S^i_{95\%\text{CL}}(\text{data})$ is the required number of signal events allowing for a 95\% C.L.\ exclusion in the presence of the measured number of events (data). The total cross section $\sigma_{\text{tot}}$ was computed from \textsc{Prospino\,2}~\cite{Beenakker:1996ch} and multiplied with NLL $K$-factors from \textsc{NLLfast}~\cite{Beenakker:2009ha}.\footnote{% For points with ${m_{\widetilde{g}}}>2500\text{\,GeV}$, we used the respective $K$-factor for ${m_{\widetilde{g}}}=2500\text{\,GeV}$, as larger values are not provided in \textsc{NLLfast}. \label{ftn:NLLgrid}} Figure \ref{fig:exclim} shows the 95\% C.L. exclusion limits -- the contours $\mu\!=\!1$ -- for the MHT and $\alpha_{\text{T}}$ analyses for first generation squarks only as well as first and second generation squarks, both for ${m_{\widetilde{g}}}/{m_{\widetilde{q}}}=2,4$. The shaded bands around the exclusion limits denote the uncertainties from scale variation, which we took to be $\mu = {m_{\widetilde{q}}}/2, 2{m_{\widetilde{q}}}$.\footnote{% We show here the scale variation at NLO\@. The information was not available at NLL accuracy for the complete parameter space considered (see footnote \ref{ftn:NLLgrid}). However, although scale uncertainties at NLL are smaller than at NLO, additional uncertainties from the parton distribution functions and $\alpha_\text{s}$ should, in principle, be taken into account. Hence, we expect the presented uncertainties to give a reasonable estimate of the over-all theoretical uncertainty.} We observe the following results: First, for both ${m_{\widetilde{g}}}/{m_{\widetilde{q}}}=2$ and $4$, the deviations in the exclusion limits derived from the efficiencies taken from $\text{T2}_\infty$ and the full supersymmetric model $\text{T2}_\mgo$ are of the order of the theoretical uncertainty on the cross section normalization for large regions in parameter space. Second, whereas for the $\alpha_{\text{T}}$ analysis the uncertainty bands overlap throughout the exclusion limit, there are deviations in the MHT analysis in the region $m_{\s\chi^0_1}\gtrsim{m_{\widetilde{q}}}/2$. The upper limit on $m_{\s\chi^0_1}$ for ${m_{\widetilde{q}}}<1\text{\,TeV}$ is very sensitive to small changes in the efficiencies. The $\alpha_{\text{T}}$ analysis is much less sensitive to the actual production mode and thus less model-dependent. Third, while for the MHT analysis the limits from $\text{T2}_\infty$ overestimate the exclusion limits for most of the parameter space, the $\alpha_{\text{T}}$ analysis $\text{T2}_\infty$ limits stay conservative over the complete parameter space. \vspace*{5mm} \begin{figure}[t] \centering \setlength{\unitlength}{1\textwidth} \begin{picture}(0.937,0.778) \put(0.52,0.42){ \put(0.0,0.025){\includegraphics[scale=1.1]{pdffig/lim_1gen_aT.pdf}} \put(0.24,0.32){\footnotesize $\alpha_{\text{T}}$, $1^{\text{st}}$\,generation} \put(0.2,0.0){\footnotesize ${m_{\widetilde{q}}}\,[\text{\,GeV}\,]$} \put(-0.04,0.14){\rotatebox{90}{\footnotesize $m_{\s\chi^0_1}\,[\text{\,GeV}\,]$}} \put(0.31,0.258){\tiny ${m_{\widetilde{g}}}/{m_{\widetilde{q}}}\!=\!2$} \put(0.2,0.18){\tiny ${m_{\widetilde{g}}}/{m_{\widetilde{q}}}\!=\!4$} \put(0.108,0.107){\scriptsize $\text{T2}_\mgo$} \put(0.108,0.074){\scriptsize $\text{T2}_\infty$ efficiency} \put(0.04,0.288){\rotatebox{37.6}{\tiny $m_{\s\chi^0_1}\!>\!{m_{\widetilde{q}}}$}} } \put(0,0.42){ \put(0.0,0.025){\includegraphics[scale=1.1]{pdffig/lim_1gen_MHT.pdf}} \put(0.22,0.32){\footnotesize MHT, $1^{\text{st}}$\,generation} \put(0.2,0.0){\footnotesize ${m_{\widetilde{q}}}\,[\text{\,GeV}\,]$} \put(-0.04,0.14){\rotatebox{90}{\footnotesize $m_{\s\chi^0_1}\,[\text{\,GeV}\,]$}} \put(0.31,0.25){\tiny ${m_{\widetilde{g}}}/{m_{\widetilde{q}}}\!=\!2$} \put(0.12,0.177){\tiny ${m_{\widetilde{g}}}/{m_{\widetilde{q}}}\!=\!4$} \put(0.108,0.103){\scriptsize $\text{T2}_\mgo$} \put(0.108,0.0697){\scriptsize $\text{T2}_\infty$ efficiency} \put(0.04,0.288){\rotatebox{37.6}{\tiny $m_{\s\chi^0_1}\!>\!{m_{\widetilde{q}}}$}} } \put(0.52,0.008){ \put(0.0,0.025){\includegraphics[scale=1.1]{pdffig/lim_2gen_aT.pdf}} \put(0.19,0.32){\footnotesize $\alpha_{\text{T}}$, $1^{\text{st}}+2^{\text{nd}}$\,generation} \put(0.2,0.0){\footnotesize ${m_{\widetilde{q}}}\,[\text{\,GeV}\,]$} \put(-0.04,0.14){\rotatebox{90}{\footnotesize $m_{\s\chi^0_1}\,[\text{\,GeV}\,]$}} \put(0.31,0.27){\tiny ${m_{\widetilde{g}}}/{m_{\widetilde{q}}}\!=\!2$} \put(0.22,0.196){\tiny ${m_{\widetilde{g}}}/{m_{\widetilde{q}}}\!=\!4$} \put(0.108,0.107){\scriptsize $\text{T2}_\mgo$} \put(0.108,0.074){\scriptsize $\text{T2}_\infty$ efficiency} \put(0.04,0.288){\rotatebox{37.6}{\tiny $m_{\s\chi^0_1}\!>\!{m_{\widetilde{q}}}$}} } \put(0,0.008){ \put(0.0,0.025){\includegraphics[scale=1.1]{pdffig/lim_2gen_MHT.pdf}} \put(0.17,0.32){\footnotesize MHT, $1^{\text{st}}+2^{\text{nd}}$\,generation} \put(0.2,0.0){\footnotesize ${m_{\widetilde{q}}}\,[\text{\,GeV}\,]$} \put(-0.04,0.14){\rotatebox{90}{\footnotesize $m_{\s\chi^0_1}\,[\text{\,GeV}\,]$}} \put(0.33,0.25){\tiny ${m_{\widetilde{g}}}/{m_{\widetilde{q}}}\!=\!2$} \put(0.05,0.147){\tiny ${m_{\widetilde{g}}}/{m_{\widetilde{q}}}\!=\!4$} \put(0.108,0.103){\scriptsize $\text{T2}_\mgo$} \put(0.108,0.0697){\scriptsize $\text{T2}_\infty$ efficiency} \put(0.04,0.288){\rotatebox{37.6}{\tiny $m_{\s\chi^0_1}\!>\!{m_{\widetilde{q}}}$}} } \end{picture} \caption{95\% C.L. exclusion limits derived from the full $\text{T2}_\mgo$ model (red, solid curves) and from the efficiencies for the $\text{T2}_\infty$ simplified model (blue, dashed curves). The shaded regions around the curves denote the uncertainties due to scale variation ($\mu={m_{\widetilde{q}}}/2,2{m_{\widetilde{q}}}$). The $\alpha_{\text{T}}$ and MHT analyses are based on an integrated luminosity of 11.7\,fb$^{-1}$ and 19.5\,fb$^{-1}$, respectively, collected at the $8\text{\,TeV}$ LHC\@. } \label{fig:exclim} \end{figure} \section{Conclusion}\label{sec:conclusion} In this study we investigated the validity of a simplified model description of squark production at the LHC\@. Concentrating on the T2 topology, $pp\to{\widetilde{q}}^{(*)}{\widetilde{q}}^{(*)}\to q q\s\chi^0_1\s\chi^0_1$, we examined the effect of a varying gluino mass and thus of varying dominant production channels. We found that the often used limiting case of a simplified model with a decoupled gluino (here denoted as $\text{T2}_\infty$) is a priori not a good description unless ${m_{\widetilde{g}}} \gg 10\,{m_{\widetilde{q}}}$. For most of the relevant parameter space for ${m_{\widetilde{g}}}<10\, {m_{\widetilde{q}}}$, squark production is dominated by the production channels $pp\to{\widetilde{q}}_\L^{}{\widetilde{q}}_\L^{},{\widetilde{q}}_{\text{R}}^{}{\widetilde{q}}_{\text{R}}^{}$; this in contrast to the case of a decoupled gluino, where only same-flavor and same-chirality squark-anti-squark production is present. We computed the signal efficiencies for two all-hadronic analyses performed by CMS: one based on missing transverse energy (MHT), and one based on the $\alpha_{\text{T}}$ variable. We found a larger sensitivity on the production channel in the MHT analysis. For the most sensitive bin in the analyses, we found relative deviations between the efficiencies from $\text{T2}_\infty$ and the dominant production mode ($pp\to{\widetilde{q}}_\L^{}{\widetilde{q}}_\L^{},{\widetilde{q}}_{\text{R}}^{}{\widetilde{q}}_{\text{R}}^{}$) of up to 70\% for the MHT analysis and up to 40\% for the $\alpha_{\text{T}}$ analysis. However, both maximal differences were found in the region of large ${m_{\widetilde{q}}}$ and small mass splittings ${m_{\widetilde{q}}}-m_{\s\chi^0_1}$, which is far beyond the exclusion limits that could be derived from the $8\text{\,TeV}$ LHC run. Hence, we found little deviation between the derived mass exclusion limits from the $\text{T2}_\infty$ simplified model efficiencies and those of the full supersymmetric model $\text{T2}_\mgo$. In particular, we found that limits derived from the $\alpha_{\text{T}}$ analysis are much less sensitive to the production mode. For the $\alpha_{\text{T}}$ analysis, the limits from $\text{T2}_\infty$ provide conservative estimates for the limits within the full model. In contrast, $\text{T2}_\infty$ tends to overestimate the limits in the case of the MHT analysis. We showed our results for ${m_{\widetilde{g}}}/{m_{\widetilde{q}}}=2,4$. For ${m_{\widetilde{g}}}/{m_{\widetilde{q}}}>4$ and ${m_{\widetilde{g}}}/{m_{\widetilde{q}}}<2$, the deviations in the efficiencies tend to become smaller. For the former case, the same-flavor and same-chirality squark-anti-squark production becomes more important, which is much less dependent on the gluino mass and hence resembles $\text{T2}_\infty$ for large ${m_{\widetilde{g}}}/{m_{\widetilde{q}}}$. For ${m_{\widetilde{g}}}/{m_{\widetilde{q}}}<2$, the production channel ${\widetilde{q}}_\L^{}{\widetilde{q}}_{\text{R}}^{}$ becomes important. However, the differences in the efficiencies between ${\widetilde{q}}_\L^{}{\widetilde{q}}_{\text{R}}^{}$ and $\text{T2}_\infty$ tend to decrease with decreasing ${m_{\widetilde{g}}}/{m_{\widetilde{q}}}$. In particular, we found smaller differences than for ${\widetilde{q}}_\L^{}{\widetilde{q}}_\L^{}$ with ${m_{\widetilde{g}}}/{m_{\widetilde{q}}}=1$. However, for ${m_{\widetilde{g}}}/{m_{\widetilde{q}}}\lesssim1$ associated squark-gluino production becomes dominant. In this part of parameter space, the gluino mass has to be taken into account as an additional parameter in the simplified model analysis. To conclude, the simplified model $\text{T2}_\infty$ is a reliable tool to interpret the current hadronic jets plus missing energy searches at the LHC in a more model-independent way, in particular for analyses based on the variable $\alpha_{\text{T}}$. Larger differences between simplified model and complete supersymmetric model interpretations could, however, arise in future LHC searches for heavier squarks, depending in detail on the experimental cuts and the parameter space probed. \section*{Acknowledgements} We would like to thank Christian Autermann, Lutz Feld and Wolfgang Waltenberger for useful discussions and suggestions. We are grateful to Wolfgang and to the Institute of High Energy Physics (HEPHY) for their hospitality during various visits to Vienna. MK is grateful to SLAC and Stanford University for their hospitality during his sabbatical stay. This work was supported by the Deutsche Forschungsgemeinschaft through the graduate school ``Particle and Astroparticle Physics in the Light of the LHC'' and through the collaborative research centre TTR9 ``Computational Particle Physics'', by the German Federal Ministry of Education and Research BMBF, and by the U.S.\ Department of Energy under contract DE-AC02-76SF00515. \addcontentsline{toc}{section}{References} \bibliographystyle{utphys}
\chapter{Introduction} \begin{abstract} In this paper, we present a new approach for analyzing gene expression data that builds on topological characteristics of time series. Our goal is to identify cell cycle regulated genes in micro array dataset. We construct a point cloud out of time series using delay coordinate embeddings. Persistent homology is utilized to analyse the topology of the point cloud for detection of periodicity. This novel technique is accurate and robust to noise, missing data points and varying sampling intervals. Our experiments using Yeast Saccharomyces cerevisiae dataset substantiate the capabilities of the proposed method. \end{abstract} \begin{IEEEkeywords} Gene expression, microarrays, topological signal analysis, periodicity detection, biomedical signal processing \end{IEEEkeywords} \section{Introduction} \blfootnote{*Research supported by National Science Foundation EEC-1160483.} Cyclic cellular regulation happens in numerous biological networks due to different regulation modes. Many genes are regulated in a periodic form coincident with the cell cycle. Finding evidence of cyclicity or periodicity in microarray cell cycle data to identify cell-cycle regulated genes is a challenging goal in gene expression time series analysis due to the following aspects of this dataset. Mostly, only a small subset of the genes show harmonically changing expression in the cycle and the corresponding time series are remarkably short including a very small number of time points that oscillate in very few number of cycles. Also, periodic gene expression signals are inherently noisy, and dominated by non-cyclic components \cite{Wichert}. Moreover, gene expression time series are not necessarily evenly sampled and the sampling frequency can vary by time. Finally, there are so many missing values in their expression profiles. Accordingly, detecting periodic gene expression time series is a very difficult problem. Therefore, development of efficient and accurate techniques for detecting cell cycle regulated genes which is robust to varying sampling rate and missing data points is of great importance. Different mathematical and statistical methods are applied to micro array time series data in order to determine periodicity \cite{Ahdesma,Johansson, Luan04,Wichert, Lu04, Liew09}. Yeast Saccharomyces cerevisiae dataset by Spellman et al \cite{Spellman} is a renowned set of gene expression time series utilized in many studies including the ones mentioned above. In Spellman's original paper, cell cycle regulated genes are identified using the combination of a Fourier transform and a correlation algorithm. A statistical test is utilized in \cite{Wichert} to identify periodically expressed genes based on the g-statistic and the false discovery rate approach is used for multiple testing. The proposed model in this study consists of a simple sinusoidal wave and additive noise while the periodicity detection is performed by searching for peaks in the periodogram. Moreover, in \cite{Liew09}, singular spectrum analysis, autoregressive based spectral estimation by signal reconstruction, and statistical hypothesis testing are used for periodic signal identification. In the existing methods, periodic models with a fixed frequency and sampling rate are used and missing data points need to be estimated as part of the pre-processing. Our proposed model for cell cycle regulated genes time series is a sinusoidal function with time varying amplitude, frequency and phase. This model makes our analysis robust to noise since disturbance changes the amplitude and period of the harmonic time series. We use time-delay coordinate embedding as a tool to construct point cloud from time series. The advantage of this technique is that the point clouds corresponding to the cell cycle regulated genes show cyclic structures even for very short time series. We can detect the presence of harmonic structures in the data by exploiting topological tools for the analysis of the delay embedding point clouds. Using topology of the point cloud, we can detect periodicity of the time series in the presence of missing data points. Persistent homology is used to detect topological holes in the delay embedding point cloud occurring on account of the periodic structure of the cell cycle regulated genes. The remainder of the paper is organized as follows: the proposed model and periodicity detection technique is presented in Section 2. Experimental results are included in Section 3. Finally Section 4 concludes the paper. \section{Method} \subsection{The Model} \label{mod} In this section, we first introduce a model for the general pattern of almost periodic time series. This model is a combination of sinusoidal functions with different frequencies, amplitudes and phases. This representation allows us to detect periodic patterns robustly since the frequency, amplitude and phase of the signal can vary with time. Therefore, we will be able to identify cyclicity in inherently noisy data where these characteristics of the signal are affected by noise. The proposed model in time domain can be expressed as: \begin{equation} \label{w} y(t)=\sum_{i=1}^n g_i(t), \end{equation} where \begin{equation}\label{g} g_i(t) = \left\lbrace \begin{array}{ccc} y_i(t) & & t_{i-1}\leq t < t_i, \\ 0 & & \text{otherwise} \end{array} \right. \end{equation} and $y_i's, i=1,2,...,n$ are defined as, \begin{equation}\label{wi} y_i(t)=A(t) \sin\left(\omega_i t+\phi_i\right), \end{equation} where $A(t)$ is a nonzero continuous amplitude function. Also, to satisfy the continuity of $\omega_i $, the phases $\phi_i$ should conform to the following condition $\phi _{i+1}=\phi _{i}+2\pi t_{i} \left( \omega_i - \omega_{i+1} \right).$ Figure \ref{model} shows two periodic genes and one non-periodic gene with their corresponding models as presented in (\ref{w})-(\ref{wi}). The first plot uses one fixed frequency i.e. $\omega_i=\omega, \forall i$ while amplitude is time varying. The second periodic gene is estimated using both time varying frequency and amplitude. The root mean square error(RMSE) between the periodic genes and their model in Figure \ref{model} are 0.20 and 0.19, respectively. On the other hand if we use the proposed model for estimation of a non-periodic gene, we get a much higher error as RMSE for the one shown in Figure \ref{model} is 0.4877. Note that the expression levels are normalized in $[-1,1]$. According to our experiments using Yeast Saccharomyces cerevisiae dataset, the average RMSE between the periodic genes and our proposed model is $8.50\%$ of the peak to peak value of the signals. On the other hand, for non-periodic genes RMSE between the model and the signal is very large, with an average of $25.93\%$ of the peak to peak value of the signals. Thus, the use of this model in the analysis techniques can address the problem of differentiating between genes with cyclic and non-cyclic patterns quite well. \begin{figure}[tb] \centering \includegraphics[width=7cm]{Figure_model.png} \caption{Top: a periodic gene and its corresponding model using fixed frequency and time varying amplitude, middle: a periodic gene and its model using time varying frequency and amplitude, bottom: a non-periodic gene and its model showing a large root mean square error} \label{model} \end{figure} \subsection{Delay Embedding for Point Cloud Construction} \begin{figure}[tb] \centering \includegraphics[width=8cm]{Picture1-5.png} \caption{Left: a sample time series, middle: its delay embedding, right: the corresponding barcode} \label{dem} \end{figure} The mathematical foundation of delay-coordinate embedding method was proposed by Packard et al. \cite{Pack} and Takens \cite{Takens} to embed a scalar time series into a higher dimensional space. The 2-dimensional delay-coordinate embedding of time series $\lbrace y_i\rbrace$ can be expressed as \begin{equation}\label{2d} Y(t)=(y(t), y(t+\tau)) \end{equation} In order for the delay-coordinate components $y(t+\tau_i),i=1,2,..,m-1$ to be independent of each other, the delay time $\tau$ needs to be chosen carefully. Roughly, if too small, then the the delay embedding is compressed along the identity line. On the other hand, for too large delays, adjacent components of the delay embedding may become irrelevant \cite{Cas91}. We examined the autocorrelation-like function (ACL) to choose a proper delay time. The empirical autocorrelation function of the signal $y(t)$ is calculated as follows \begin{equation}\label{ac} R_{yy}(t)=\sum_k y(k+t)y(k) \end{equation} Note that for time varying non-stationary signals, a strict autocorrelation function cannot be defined and used for neither establishing periodicity not for detecting the frequencies. However, we use Equation (\ref{ac}) as an autocorrelation-like function strictly for delay selection. Clearly, peaks in the autocorrelation-like function mark delay times at which the signal is comparatively highly correlated with itself. According to experimental results, the appropriate interval for choosing delay time to best obtain informative delay embedding of signal $x(t)$ can be expressed as \begin{equation}\label{td} t_{c1}<\tau<t_{c2}, \end{equation} where $t_{c1}$ and $t_{c2}$ are the first and second critical points of the autocorrelation-like function $R_{yy}(t)$. We have proved in \cite{me14} that the delay-coordinate embedding $Y(t)$ of $y(t)$ as described in (\ref{w})-(\ref{wi}) is a set of concentric ellipses, with varying radii and side lengths of the circumscribed squares around them. It is also shown that $Y(t)$ obtained using the appropriate delay as expressed in (\ref{td}) always has a topological hole. According to the discussions and results presented in \ref{mod}, the model $Y(t)$ approximates the time series of the cyclic genes pretty well. Therefore, there is always a topological hole inside the delay embedding of periodic genes as well. This hole will be used in microarray data in order to detect the periodicity in cell cycle regulated genes. \begin{figure}[tb] \centering \includegraphics[width=4cm]{Picture2.png} \caption{Illustration of persistant homolgy algorithm and barcode construction } \label{PH} \end{figure} \subsection{Persistent Homology} \label{PHAlg} The persistent homology technique is utilized as a tool to identify the hole inside delay embedding of periodic genes using a filtration of the point clouds. This method characterizes the topology of the point clouds using a collection of betti intervals, or persistence intervals which are encapsulated in a structure called persistent barcode. Given a point cloud of $n$ points $v_1, v_2,...,v_n$, persistent homology constructs a barcode using the algorithm described below and illustrated in Figure \ref{PH}. We create balls of radius $r$ around each point in the point cloud $v_i$ and denote it by $B_r(v_i)$. The radius $r$ is increased until the balls intersect. For all $0 < i,j < n, i \neq j$, we add a line $[v_1, v_2]$ if and only if $B_r(v_i) \cap B_r(v_i) \neq \emptyset$. While lines are forming and connecting the points, every time the skeleton of a triangle shapes we fill the triangle. To construct the barcode we compute the cycles at each radius value $r$ and represent each cycle(hole) as a point in a 2 dimensional plot where its x-coordinate is $r$. The height of each point is arbitrary but it stays the same for each hole by varying $r$. As we proceed in our filtration sequence by increasing the radius $r$, each Betti interval or bar in the barcode $[r_b,r_d]$ can be associated to a hole, which appears at radius $r_b$ (birth) and ``closes" at $r_d$ (death) \cite{Carlsson08}. Length of each bar in the barcode is a representation of the size of the corresponding hole. Since the point cloud constructed from the time series of genes with periodic structure always has one dominant hole, the length of the longest bar in the barcode represents that hole. On the other hand, the point cloud of non periodic genes does not contain any dominant hole and only has a few small holes disappearing shortly when $r$ grows. Therefore, barcodes of non-cyclic genes include some short bars that die soon. Accordingly, we can identify cell cycle regulated genes using a threshold on the length of the longest bar in their persistent barcodes. This approach is robust to missing data points since a small subset of the data points in the point cloud preserves the topology of the data set. The use of this approach in the micro array data can address the problem of identifying cell cycle regulated genes quite well. \section{Results} The analysis is performed on the yeast Saccharomyces cerevisiae experiment dataset available at http://genome-www.stanford.edu/cellcycle/. This dataset consists of DNA microarrays and samples from yeast cultures synchronized using three different techniques, including $\alpha$ factor arrest (alpha), elutriation synchronization and temperature arrests cdc 15 and cdc28. This dataset contains information about 6178 genes. The $\alpha$ factor arrest dataset includes 18 time points with a sampling period of 7 mins. The cdc 15 data is sampled at 20 mins for the first 4 data point, 10 mins for the next 18 time points and again 20 mins for the last two points. The measurements in cdc28 experiment are done at intervals of 10 mins from 0 to 160 min with a total of 17 time points. The sampling period of elutriation experiment is 30 mins and it includes a total of 14 points from 0 to 390 min. To date 104 cell cycle regulated genes have been identified in this daatset using traditional techniques \cite{Spellman}. We normalize the amplitudes of all the time series between -1 and 1. In the delay embedding construction, the delays are chosen using Equation (\ref{td}) for the autocorrelation like function. The selected delays for alpha, cdc15, cdc28 and elutriation dataset are 2, 3, 2 and 3 samples, respectively. The delay embedding of all time series are then constructed using equation (\ref{2d}). Persitant homology algorithm as described in \ref{PHAlg} is utilizes using javaplex package for Matlab \cite{javaPlex} to build the corresponding barcodes. Figure \ref{DEm} shows time series of one periodic and one non-periodic gene in alpha dataset in the first row, while their delay embeddings are represented in second row. Clearly, the delay embedding of the cyclic gene contains a dominant hole while the point cloud of the gene with non-periodic pattern does not have any cyclic structure and includes no visible hole. The last row depicts the barcodes showing a long bar for the periodic gene and a few very short bars for the non-cyclic gene data. This experiment is performed on the whole dataset and the length of the longest bar is calculated for each gene. The results for alpha dataset is shown in Figure \ref{Thresh}. Four different thresholds are chosen experimentally and the genes with the length of the longest bar greater than the threshold are considered cell cycle regulated genes. The false negative and false positive numbers for these thresholds (0.16, 0.2, 0.24, 0.28) are tabulated in Table I. Using alpha, elutriation, cdc28 and cdc15 dataset, we were able to identify $85.58\%$, $87.01\%$, $76.33\%$ and $71.60\%$ of the total of 104 cell cycle regulated genes, respectively. \begin{figure}[tb] \centering \includegraphics[width=1\linewidth]{Figure3.png} \caption{Top row: a periodic (left) and a non-periodic (right) time series. Middle row: their corresponding delay embeddings. Bottom row: their corresponding dimension 1 barcode} \label{DEm} \end{figure} \begin{figure}[tb] \centering \includegraphics[width=0.8\linewidth]{alpha_L.png} \caption{The length of the longest bar in the barcodes of all genes in alpha dataset} \label{Thresh} \end{figure} \begin{table}[tb] \begin{center} \caption{Experimental results for alpha dataset.} \label{Table1Label} \begin{tabular}{|c|c|c|} \hline Thershold & False Negative & False Positive\\ \hline 0.16 & 15 ($14.42\%$) & 47 \\ \hline 0.2 & 15 ($14.42\%$) & 39 \\ \hline 0.24 & 18 ($17.31\%$) & 52 \\ \hline 0.28 & 18 ($17.31\%$) & 71 \\ \hline \end{tabular} \end{center} \end{table} \section{Conclusion} In this study, we proposed topological analysis of delay-coordinate embedding to identify cell cycle regulated genes in microarray data. A new model for cyclic gene time series is introduced as a piecewise sinusoidal wave with time varying frequency, amplitude and phase. We described how to apply peristant homology algorithm to delay embedding point cloud of gene expression data. The presented approach were motivated by the challenges in the analysis of microarray cell cycle data including very short time series, many missing data points, intrinsic noise and time varying sampling rates. We applied our methods to the real expression data of Spellman et al. \cite{Spellman} and used periodic genes identified by traditional techniques for comparison of the results. \bibliographystyle{ieeetr}
\section{Introduction} \vspace{-2mm} Lepton flavor violation (LFV) is the clearest signal for physics beyond the Standard Model (SM)~\cite{Kuno:1999jp}, and extensive searches for LFV have been made since the muon was found~\cite{Brooks:1999pu, Adam:2013mnn, Bertl:2006up, Bellgardt:1987du}. Though a lot of efforts have been made, we have not found any LFV signals with charged leptons. LFV had, however, been found in neutrino oscillation \cite{Fukuda:1998mi,Abe:2013hdq} and it indeed requires us to extend the SM so that physics beyond the SM must include LFV. This fact also gives us a strong motivation to search for charged lepton flavor violation (cLFV). Along this line new experiments to search for cLFV will start soon. COMET \cite{Cui:2009zz,Kuno:2013mha} and DeeMe \cite{Natori:2014yba} will launch within a few years and search $\mu$-$e$ conversion. In these experiments, first, muons are trapped by target nucleus, then, if cLFV exists, it converts into an electron. If COMET/DeeMe observe the $\mu$-$e$ conversion, then with what kind of new physics should we interpret it? Now it is worth considering since we are in-between two kinds of cLFV experiments with muon. For these several decades, theories with supersymmetric extension have been most studied. These theories include a source of LFV. It is realized by the fact that the scalar partner of charged leptons have a different flavor basis from that of charged leptons. In addition, R-parity is often imposed on this class of the theory\cite{Hisano:1995cp,Sato:2000ff}. With it, $\mu\rightarrow e\gamma$ process has the largest branching ratio among the three cLFV processes. This occurs through the dipole process depicted and the other two, $\mu-e$ conversion and $\mu\rightarrow 3e$, are realized by attaching a quark line and an electron line at the end of the photon line respectively, giving an O($\alpha$) suppression. Those branching ratios must be smaller than that of $\mu\rightarrow e\gamma$. At this moment, however, the upper bounds for those branching ratios are almost same each other. It means if COMET/DeeMe observe the $\mu-e$ conversion, we have to discard this scenario. It is, however, possible to find a theory easily in which COMET/DeeMe find cLFV first. To see this we first note that the $\mu\rightarrow e \gamma$ process occurs only at loop level due to the gauge invariance, while other two can occur as a tree process. Therefore in this case we have to consider a theory in which the $\mu-e$ conversion process occurs as tree process. In other words we have to assume a particle which violate muon and electron number. Since $\mu - e$ conversion occurs in a nucleus, it also couples with quarks with flavor conservation. Furthermore it is better to assume that it does not couple with two electrons as we have not observed $\mu\rightarrow 3e$. In this paper we consider the case that COMET/DeeMe indeed observe the cLFV process, while all the other experiments will not observe anything new at that time. With this situation, we need to understand how to confirm the cLFV in other experiments. Unfortunately in this case other new physics signals are expected to be quite few, since the magnitude of the cLFV interaction is so small due to its tiny branching ratio. Therefore it is very important to simulate now how to confirm the COMET signal and the new physics. As a benchmark case we study supersymmetric models without R parity~\cite{deGouvea:2000cf} . In this kind of theory the scalar lepton mediates $\mu \leftrightarrow e$ flavor violation. \section{RPV interaction and our scenario} \label{Sec:interaction} \vspace{-2mm} In general the supersymmetric gauge invariant superpotential contains the R-parity violating terms~\cite{Weinberg:1981wj, Sakai:1981pk, Hall:1983id}, \begin{equation} \begin{split} \mathcal{W}_\text{RPV} = \lambda_{ijk} L_i L_j E_k^c + \lambda'_{ijk} L_i Q_j D_k^c + \lambda''_{ijk} U_i^c D_j^c D_k^c, \label{Eq:RPV_SP} \end{split} \end{equation} where $E_i^c$, $U_i^c$ and $D_i^c$ are $SU(2)_L$ singlet superfields, and $L_i$ and $Q_i$ are $SU(2)_L$ doublet superfields. Indices $i$, $j$, and $k$ represent the generations. We take $\lambda_{ijk} = - \lambda_{jik}$ and $\lambda''_{ijk} = - \lambda''_{ikj}$. First two terms include lepton number violation, and the last term includes baryon number violation. Since some combinations of them accelerate proton decay, we omit the last term. Our interesting situation is that only $\mu$-$e$ conversion is discovered, and other cLFV processes will never be observed. The situation is realized under the following 3 setting on the RPV interaction: \begin{enumerate} \item only the third generation slepton contributes to the RPV interactions \item for quarks, flavor diagonal components are much larger than that of off-diagonal components, i.e., CKM-like matrix, $\lambda'_{ijj} \gg \lambda'_{ijk} (j \neq k)$ \item the generation between left-handed and right-handed leptons are different, $\lambda_{ijk} (i \neq k \text{ and } j \neq k)$. \end{enumerate} The setting-1 is naturally realized by the RG evolved SUSY spectrum with universal soft masses at the GUT scale. For the simplicity, we decouple SUSY particles except for the third generation sleptons. The setting-2 is also realized in most cases unless we introduce additional sources of flavor violations. The setting-3 is artificially introduced to realize the interesting situation (see Introduction). Under the settings, the Lagrangian from the superpotential~\eqref{Eq:RPV_SP} is reduced as follows, \begin{equation} \begin{split} & \mathcal{L}_\text{RPV} = \mathcal{L}_{\lambda} + \mathcal{L}_{\lambda'}, \\& \mathcal{L}_\lambda = 2 \bigl[ \lambda_{312} \tilde \nu_{\tau L} \overline{\mu} P_L e + \lambda_{321} \tilde \nu_{\tau L} \overline{e} P_L \mu + \lambda_{132} \tilde \tau_L \overline{\mu} P_L \nu_e + \lambda_{231} \tilde \tau_L \overline{e} P_L \nu_\mu \\& \hspace{10mm} + \lambda_{123} \tilde \tau_R^* \overline{(\nu_{eL})^c} P_L \mu + \lambda_{213} \tilde \tau_R^* \overline{(\nu_{\mu L})^c} P_L e \bigr] + \text{h.c.}, \\& \mathcal{L}_{\lambda'} = \bigl[ \lambda'_{311} \bigl( \tilde \nu_{\tau L} \overline{d} P_L d - \tilde \tau_L \overline{d} P_L u \bigr) + \lambda'_{322} \bigl( \tilde \nu_{\tau L} \overline{s} P_L s - \tilde \tau_L \overline{s} P_L c \bigr) \bigr] + \text{h.c.}. \label{Eq:RPV_L2} \end{split} \end{equation} Some kind of processes described by the Lagrangian \eqref{Eq:RPV_L2} strongly depend on the values of $\lambda'_{311}$ and $\lambda'_{322}$. In order to clarify the dependence and to discuss the discrimination of each other, we study following three cases in our paper~\cite{RPV}: [case-I] $\lambda'_{311} \neq 0$ and $\lambda'_{322} = 0$, [case-I\hspace{-1pt}I] $\lambda'_{311} = 0$ and $\lambda'_{322} \neq 0$, and [case-I\hspace{-1pt}I\hspace{-1pt}I] $\lambda'_{311} \neq 0$ and $\lambda'_{322} \neq 0$. In this talk, we focus on the case-I. \section{Exotic processes in our scenario} \label{Sec:Obs} In the scenario we have five types of exotic processes: \\[2mm] ~~~~~ 1 ~~ $\mu$-$e$ conversion in a nucleus ($\mu^- N \to e^- N$) \\ ~~~~~ 2 ~~ $\mu^- e^+$ production at LHC ($pp \to \mu^- e^+$) \\ ~~~~~ 3 ~~ dijet production at LHC ($pp \to jj$) \\ ~~~~~ 4 ~~ non-standard interaction (NSI) of neutrinos \\ ~~~~~ 5 ~~ muonium conversion ($\mu^+e^- \to \mu^-e^+$) \\[2mm] In the situation that the $\mu$-$e$ conversion is discovered while other cLFV signals will never be found, we discuss the possibility whether we can confirm the $\mu$-$e$ conversion signal with the five types processes or not. Details of each process and the formulation of their reaction rates are given in our paper~\cite{RPV}. Note that in our scenario other muon cLFV processes ($\mu \to e \gamma$, $\mu \to 3e$, $\mu^- e^- \to e^- e^-$ in muonic atom~\cite{Koike:2010xr}, and so on) occur at two-loop level. At one glance the tau sneutrino can connect with the photon via d-quark loop. The contribution of the loop of the diagram is \begin{eqnarray} \lambda'\left(-\frac{1}{3}\right) e \frac{m_d q_\mu}{8\pi^2} \int_0^1 dx(1-2x)\log(m_d^2-(x-x^2)q^2)\ \propto q^2q^\mu, \end{eqnarray} where $q$ is the momentum of the photon. The contribution to cLFV is, therefore vanish with on-shell photon ($q^2=0$) for $\mu\rightarrow e\gamma$ and with $\bar e \gamma_\mu e$ attached for $\mu\rightarrow 3e$ due to gauge symmetry($q^\mu\bar e\gamma_\mu e=0$). Thus these processes occur at two-loop level. Furthermore these processes are extremely suppressed by higher order couplings, gauge invariance, and so on. Therefore we do not study these processes. \section{Numerical result} \label{Sec:result} \vspace{-2mm} \begin{table}[t] \caption{Current and future experimental limits on the $\mu$-$e$ conversion branching ratio and the upper limits on $\lambda' \lambda$ corresponding to each experimental limit. } \vspace{-5mm} \hspace{-3mm} \small{ {\renewcommand\arraystretch{1.7} \begin{tabular}{llllll} \hline Experiment & BR limit & Limit on $\lambda'_{311} \lambda$ (case-I) & Limit on $\lambda'_{322} \lambda$ (case-I\hspace{-1pt}I) & Limit on $\lambda' \lambda$ (case-I\hspace{-1pt}I\hspace{-1pt}I) \\ \hline \hline SINDRUM & $7 \times 10^{-13}$~\cite{Bertl:2006up} & $1.633 \times 10^{-7} \Bigl( \dfrac{m_{\tilde \nu_\tau}}{1\text{TeV}} \Bigr)^2$ & $3.170 \times 10^{-7} \Bigl( \dfrac{m_{\tilde \nu_\tau}}{1\text{TeV}} \Bigr)^2$ & $1.072 \times 10^{-7} \Bigl( \dfrac{m_{\tilde \nu_\tau}}{1\text{TeV}} \Bigr)^2$ \\[0.5mm] DeeMe & $5 \times 10^{-15}$~\cite{Natori:2014yba} & $ 1.550 \times 10^{-8} \Bigl( \dfrac{m_{\tilde \nu_\tau}}{1\text{TeV}} \Bigr)^2$ & $ 2.915 \times 10^{-8} \Bigl( \dfrac{m_{\tilde \nu_\tau}}{1\text{TeV}} \Bigr)^2$ & $ 1.012 \times 10^{-8} \Bigl( \dfrac{m_{\tilde \nu_\tau}}{1\text{TeV}} \Bigr)^2$ \\[0.5mm] COMET-I & $7 \times 10^{-15}$~\cite{Kuno:2013mha} & $1.830 \times 10^{-8} \Bigl( \dfrac{m_{\tilde \nu_\tau}}{1\text{TeV}} \Bigr)^2$ & $3.504 \times 10^{-8} \Bigl( \dfrac{m_{\tilde \nu_\tau}}{1\text{TeV}} \Bigr)^2$ & $1.196 \times 10^{-8} \Bigl( \dfrac{m_{\tilde \nu_\tau}}{1\text{TeV}} \Bigr)^2$ \\[0.5mm] COMET-I\hspace{-1pt}I &$3 \times 10^{-17}$~\cite{Kuno:2013mha} & $1.198 \times 10^{-9} \Bigl( \dfrac{m_{\tilde \nu_\tau}}{1\text{TeV}} \Bigr)^2$ & $2.294 \times 10^{-9} \Bigl( \dfrac{m_{\tilde \nu_\tau}}{1\text{TeV}} \Bigr)^2$ & $7.827 \times 10^{-10} \Bigl( \dfrac{m_{\tilde \nu_\tau}}{1\text{TeV}} \Bigr)^2$ \\[0.5mm] PRISM & $7 \times 10^{-19}$~\cite{Kuno:2013mha} & $1.830 \times 10^{-10} \Bigl( \dfrac{m_{\tilde \nu_\tau}}{1\text{TeV}} \Bigr)^2$ & $3.504 \times 10^{-10} \Bigl( \dfrac{m_{\tilde \nu_\tau}}{1\text{TeV}} \Bigr)^2$ & $1.196 \times 10^{-10} \Bigl( \dfrac{m_{\tilde \nu_\tau}}{1\text{TeV}} \Bigr)^2$ \\ \hline \end{tabular} } } \label{Tab:mue_conv} \end{table} We are now in a position to show numerical results. Table~\ref{Tab:mue_conv} shows the current experimental limit and the future single event sensitivity for $\mu$-$e$ conversion process, and shows the upper limits on the combination of the RPV couplings, $\lambda' \lambda$, corresponding to the limit and the sensitivities in each experiment. In the calculation of the upper limits, we take Au, Si, and Al for target nucleus of SINDRUM-I\hspace{-1pt}I, DeeMe, and other experiments, respectively. $\mu$-$e$ conversion search is a reliable probe to both the RPV couplings and tau sneutrino mass. The current experimental limit puts strict limit on the RPV couplings, $\lambda' \lambda \lesssim 10^{-7}$ for $m_{\tilde \nu_\tau} = 1\text{TeV}$ and $\lambda' \lambda \lesssim 10^{-5}$ for $m_{\tilde \nu_\tau} = 3\text{TeV}$, respectively. In near future, the accessible RPV couplings will be extended by more than 3 orders of current limits, $\lambda' \lambda \simeq 10^{-10}$ for $m_{\tilde \nu_\tau} = 1\text{TeV}$ and $\lambda' \lambda \simeq 10^{-8}$ for $m_{\tilde \nu_\tau} = 3\text{TeV}$, respectively. The $\mu$-$e$ conversion process is one of the clear signatures for the RPV scenario, but it is not the sufficient evidence of the scenario. We must check the correlations among the reaction rates of $\mu$-$e$ conversion process, the cross sections of $pp \to \mu^- e^+$ and $pp \to jj$, and so on in order to discriminate the case-I, -I\hspace{-1pt}I, and -I\hspace{-1pt}I\hspace{-1pt}I each other and to confirm the RPV scenario. \begin{figure}[t!] \hspace{-7mm} \begin{tabular}{cc} \subfigure[$m_{\tilde \nu_\tau} = 1$TeV. $\sqrt{s}= 14$TeV. ]{ \includegraphics[scale=0.61]{contour_case1_1TeV_14TeV.eps} \label{left1} } & \hspace{-10mm} \subfigure[$m_{\tilde \nu_\tau} = 1$TeV. $\sqrt{s}= 100$TeV. ]{ \includegraphics[scale=0.61]{contour_case1_1TeV_100TeV.eps} \label{right1} } \\ \subfigure[$m_{\tilde \nu_\tau} = 3$TeV. $\sqrt{s}= 14$TeV. ]{ \includegraphics[scale=0.61]{contour_case1_3TeV_14TeV.eps} \label{left2} } & \hspace{-10mm} \subfigure[$m_{\tilde \nu_\tau} = 3$TeV. $\sqrt{s}= 100$TeV. ]{ \includegraphics[scale=0.61]{contour_case1_3TeV_100TeV.eps} \label{right2} } \\ \end{tabular} \caption{Contour plot of $\sigma(pp \to \mu^- e^+)$, $\sigma(pp \to dijet)$, and BR($\mu^- N \to e^- N$) in the case-I for (a) $m_{\tilde \nu_\tau} = 1$TeV and $\sqrt{s}=14$TeV (b) $m_{\tilde \nu_\tau} = 1$TeV and $\sqrt{s}=100$TeV (c) $m_{\tilde \nu_\tau} = 3$TeV and $\sqrt{s}=14$TeV (d) $m_{\tilde \nu_\tau} = 3$TeV and $\sqrt{s}=100$TeV. Light shaded region is excluded by the $\mu$-$e$ conversion search~\cite{Bertl:2006up}, and dark shaded band is excluded region by the $M$-$\bar M$ conversion search~\cite{Willmann:1998gd}.} \label{Fig:cont_I} \end{figure} \begin{figure}[t!] \hspace{-6mm} \begin{tabular}{cc} \subfigure[$\text{N}=\text{Si}$ and $\sqrt{s}= 14$TeV.]{ \includegraphics[scale=0.58]{case1_Si_14TeV.eps} \label{case1_a} } & \hspace{-14mm} \subfigure[$\text{N}=\text{Si}$ and $\sqrt{s}= 100$TeV.]{ \includegraphics[scale=0.58]{case1_Si_100TeV.eps} \label{case1_b} } \\[-4.5mm] \subfigure[$\text{N}=\text{Al}$ and $\sqrt{s}= 14$TeV.]{ \includegraphics[scale=0.58]{case1_Al_14TeV.eps} \label{case1_c} } & \hspace{-14mm} \subfigure[$\text{N}=\text{Al}$ and $\sqrt{s}= 100$TeV.]{ \includegraphics[scale=0.58]{case1_Al_100TeV.eps} \label{case1_d} } \\ \end{tabular} \caption{$\sigma (pp \to \mu^- e^+)$ as a function of $\text{BR} (\mu^- N \to e^- N)$ for each $\sigma (pp \to jj)$ in the case-I. $\sigma (pp \to jj)$ are attached on each line. Results for $m_{\tilde \nu_\tau} = 1\text{TeV}$ ($m_{\tilde \nu_\tau} = 3\text{TeV}$) are given by dot-dashed line (dotted line). Shaded region in each panel is the excluded region by the SINDRUM-I\hspace{-1pt}I experiment. Left panels show the results for the collision energy $\sqrt{s} = 14\text{TeV}$, and right panels show the results for $\sqrt{s} = 100\text{TeV}$. We take Si [(a) and (b)], and Al [(c) and (d)] for the target nucleus of $\mu$-$e$ conversion process. } \label{Fig:sigma_vs_BR_I_1} \end{figure} The parameter dependence of $\sigma(pp \to \mu^- e^+)$, $\sigma(pp \to jj)$, and $\text{BR}(\mu^- N \to e^- N)$ in the case-I are depicted in Fig.~\ref{Fig:cont_I}. Dashed and dot-dashed lines are contours of $\sigma(pp \to \mu^- e^+)$ and $\sigma(pp \to jj)$ at $\sqrt{s}=14$TeV (left panels) and $\sqrt{s} =100$TeV (right panels), respectively. Solid lines are contours of $\text{BR}(\mu^- \text{Al} \to e^- \text{Al})$, which are translated from the single event sensitivities of each experiments (see Table~\ref{Tab:mue_conv}). Light shaded region is excluded by the $\mu$-$e$ conversion search at SINDRUM-I\hspace{-1pt}I~\cite{Bertl:2006up}, and dark shaded band is excluded region by the $M$-$\bar M$ conversion search experiment~\cite{Willmann:1998gd}. For simplicity, we take the couplings universally in leptonic RPV sector: $\lambda \equiv \lambda_{312} = \lambda_{321} = -\lambda_{132} = -\lambda_{231}$. Figure~\ref{Fig:cont_I} displays the strong potential of $\mu$-$e$ conversion search to explore the RPV scenarios. The PRISM experiment will cover almost parameter space wherein the LHC experiment can survey. In the parameter range between the SINDRUM-I\hspace{-1pt}I limit and the PRISM reach, combining the measurement results of $\sigma(pp \to \mu^- e^+)$, $\sigma(pp \to jj)$, and $\text{BR}(\mu^- \text{Al} \to e^- \text{Al})$, the RPV couplings and the tau sneutrino mass will be precisely determined. Figure~\ref{Fig:sigma_vs_BR_I_1} shows $\sigma(pp \to \mu \bar e)$ as a function of $\text{BR} (\mu + N \to e + N)$ in the case-I. Candidate materials for the target of $\mu$-$e$ conversion search are silicon (Si) at the DeeMe, and are aluminum (Al) at the COMET, Mu2e, and PRISM. Vertical dotted lines show the experimental reach of DeeMe 1-year running (DeeMe(1yr)), DeeMe 4-years running (DeeMe(4yrs)), COMET phase-I (COMET-I), COMET phase-I\hspace{-1pt}I (COMET-I\hspace{-1pt}I), and PRISM (PRISM). Shaded regions are the excluded region by the SINDRUM-I\hspace{-1pt}I~\cite{Bertl:2006up}, which are translated into the limit for each nucleus from that for Au. Each line corresponds to the dijet production cross section at the LHC, $\sigma(pp \to jj)$, at $\sqrt{s}=14\text{TeV}$ (left panels) and at $\sqrt{s}=100\text{TeV}$ (right panels), respectively. For simplicity, we take universal RPV coupling, $\lambda \equiv \lambda_{312} = \lambda_{321} = -\lambda_{132} = -\lambda_{231}$. Figure~\ref{Fig:sigma_vs_BR_I_1} shows the clear correlations among $\sigma (pp \to \mu^- e^+)$, $\sigma(pp \to jj)$, and $\text{BR} (\mu^- N \to e^- N)$. Checking the correlations makes possible to distinguish the RPV scenario and other new physics scenarios. \section{Summary and discussion} \label{Sec:summary} \vspace{-2.5mm} We have studied a supersymmetric standard model without R parity as a benchmark case that COMET/DeeMe observe $\mu - e$ conversion prior to all the other experiments observing new physics. In this case with the assumption that only the third generation sleptons contribute to such a process, we need to assume that $\{\lambda'_{311} {\rm \hspace{1.2mm} and/or} \hspace{1.2mm} \lambda'_{322}\} \times \{\lambda_{312}{\rm{ \hspace{1.2mm} and/or \hspace{1.2mm} }} \lambda_{321}\}$ must be sufficiently large. Though other combinations of coupling constants can lead a significant $\mu-e$ conversion process, only those are considered here. This is because in most of scenarios in the supersymmetric theory, the third generation of the scalar lepton has the lightest mass. With these assumptions, we calculated the effects on future experiments. First we considered the sensitivity of the future $\mu - e$ conversion experiments on the couplings and the masses. Then with the sensitivity kept into mind we estimated the reach to the couplings by calculating the cross section of $pp \rightarrow \mu^- e^+$ and $pp \rightarrow jj$ as a function of the slepton masses and the couplings. To have a signal of $\mu^- e^+$ both the coupling $\lambda'$ and $\lambda$ must be large and hence there are lower bounds for them while to observe dijet event via the slepton only the coupling $\lambda'$ must be large and hence there is a lower bound on it. In all cases we have a chance to get confirmation of $\mu - e $ conversion in LHC indirectly. In addition, we put a bound on the couplings by comparing both modes. Finally we considered muonium conversion. If $\lambda'$ is very small we cannot expect a signal from LHC. In this case at least one of $\lambda_{312}$ and $\lambda_{321}$ must be very large and if it is lucky, that is both of them are very large we can expect muonium conversion. There are other opportunities to check the result on $\mu -e$ conversion. For example we can distinguish $\lambda_{312}$ and $\lambda_{321}$ in linear collider with polarized beam. We can also expect the signal $p e^- \to p \mu^-$ in LHeC. It is however beyond the scope of this paper to estimate their sensitivities and we leave them in future work~\cite{future}. \section*{Acknowledgments} \vspace{-2.5mm} This work was supported in part by the Grant-in-Aid for the Ministry of Education, Culture, Sports, Science, and Technology, Government of Japan, No. 25003345 (M.Y.). \vspace{-2mm}
\section{Introduction} Since the seminal paper \cite{deWit:1982ig} in which the first gauged maximal supergravity was constructed with gauge group ${\rm SO}(8)$, much work has been done to study the vacua of this model and to construct new gauged maximal supergravities. Certain vacua of the original ${\rm SO}(8)$-gauged model\footnote{See \cite{Warner:1983vz} for early results in the search for vacua of the original theory and \cite{Fischbacher:2008zu} for a recent study.}, like the anti-de Sitter (AdS) vacuum with $\mathcal{N}=8$ residual supersymmetry, were put in correspondence with compactifications of $D=11$ supergravity on a seven-dimensional sphere or on warped/stretched versions of a seven-sphere, possibly with torsion. Non-compact and even non-semisimple gaugings, defined by groups of the form ${\rm CSO}(p,q,r)$, $p+q+r=8$, were first constructed in \cite{Hull:1984qz} and their de Sitter vacua put in correspondence with reductions on non-compact spaces with negative curvature \cite{Hull:1988jw}. Flat-gaugings in $D=4$ describing Scherk-Schwarz reductions of maximal $D=5$ supergravity and yielding no-scale models, were first constructed in \cite{Andrianopoli:2002mf}.\par A new formulation of gauged extended supergravities in terms of the so called \emph{embedding tensor} \cite{deWit:2002vt,deWit:2005ub,deWit:2007mt}, has opened the way for a more systematic analysis of the possible gaugings and their vacuum structure. All possible choices of gauge groups in the maximal supergravity are encoded in a single object $\Theta_M{}^\alpha$ (the embedding tensor), which defines the embedding of the gauge algebra inside the algebra $\mathfrak{e}_{7(7)}$ of the on-shell global symmetry group ${\rm E}_{7(7)}$ of the ungauged theory. This object is formally ${\rm E}_{7(7)}$--covariant and is constrained, by linear and quadratic conditions originating from the requirement of supersymmetry and gauge invariance, to belong to certain orbits of the ${\bf 912}$ representation. An interesting feature of this formulation is that the field equations and Bianchi identities of the gauged model are formally ${\rm E}_{7(7)}$-covariant if the fields are transformed together with the embedding tensor. In other words there is a mapping (or duality) between gauged theories defined by embedding tensors that are related by ${\rm E}_{7(7)}$ transformations. Such mapping should encode the effect of string/M-theory dualities on flux compactifications. In particular the scalar potential $V(\Theta,\phi)$ is a quadratic function of $\Theta_M{}^\alpha$ and is invariant under the simultaneous action of ${\rm E}_{7(7)}$ on the $70$ scalar fields $\phi=(\phi^{ijkl})$ of the model and the embedding tensor: \begin{equation} \forall g\in {\rm E}_{7(7)}\,\,;\,\,\,\,V(\Theta,\phi)=V(g\star \Theta,g\star\phi)\,, \end{equation} where $g\star$ denotes the generic action of a group element $g$ on the scalars (non-linear action) and on $\Theta_M{}^\alpha$ (linear action). The above property and the homogeneity of the scalar manifold has motivated what has been dubbed as the ``going-to-the-origin'' approach for the study of vacua of gauged supergravities \cite{Dibitetto:2011gm,DallAgata:2011aa}: Any vacuum of a given gauged model can be mapped into the origin of the scalar manifold\footnote{By origin we mean the point in the scalar coset manifold ${\rm E}_{7(7)}/{\rm SU}(8)$ at which all scalar fields $\phi^{ijkl}$ vanish and is thus manifestly ${\rm SU}(8)$-invariant.} by means of a suitable ${\rm E}_{7(7)}$-transformation, provided the embedding tensor is transformed accordingly. This means that the vacua of gauged maximal supergravity can be systematically studied by restricting to the origin of the manifold so that the extremization condition on $V$ becomes another condition on $\Theta_M{}^\alpha$ only. In this way, one can search for vacua with particular properties without committing to a particular gauge group, i.e.\ while simultaneously scanning through all possible gaugings.\par In \cite{DallAgata:2012bb} a new family of ${\rm SO}(8)$-gauged maximal supergravities was constructed by exploiting the freedom in the original choice of the symplectic frame defining the electric and magnetic gauge fields. These models were obtained as a deformation of the original de Wit and Nicolai model, parametrized by an angle $\omega$. They all exhibit an $\mathcal{N}=8$ vacuum at the origin. Their spectrum is identical while the $\omega$ parameter only affects the higher-order interactions. Similar generalizations of non-compact gaugings were studied in \cite{DallAgata:2012sx,DallAgata:2014ita}. Adopting the ``going-to-the-origin'' approach, the authors of \cite{Borghese:2012qm,Borghese:2012zs,Borghese:2013dja} systematically searched for vacua with certain residual symmetries and found several vacua of the new $\omega$-deformed models. An interesting feature observed in all the above works, is that the $\omega$-deformed models in general exhibit a much richer vacuum structure that the original $\omega=0$ models from \cite{deWit:1982ig,Hull:1984qz}. In other words, many vacua of these theories disappear in the limit $\omega\rightarrow 0$.\par In the present paper we start a systematic analysis of vacua of maximal supergravity with a minimal amount of residual supersymmetry. We focus on AdS vacua preserving $\mathcal{N}>2$ supersymmetries by implementing the supersymmetry conditions (Killing spinor equations) directly on the embedding tensor (with the scalar fields fixed at the origin). Our (computer aided) analysis is systematic and we find, aside from the known $\mathcal{N}=8$ vacua, only two other classes of solutions with residual supersymmetry $\mathcal{N}=4$ and $\mathcal{N}=3$, respectively. \emph{These are, to our knowledge, the first AdS vacua of maximal supergravity with residual $\mathcal{N}>2$ supersymmetry, aside from the $\mathcal{N}=8$ ones.} We can exclude, by general argument, solutions with $8>\mathcal{N}>4$. Each class of the newly found vacua is parametrized by an angle $\varphi$ and, depending on its values, the corresponding vacua are embedded in different ($\omega$-deformed) ${\rm CSO}(p,q,r)$ models. In particular the $\mathcal{N}=4$ vacua, depending on $\varphi$, belong to gaugings of the form ${\rm SO}(1,7)$ and $[{\rm SO}(1,1)\times {\rm SO}(6)]\ltimes T^{12}$, while $\mathcal{N}=3$ vacua to models with gauge group ${\rm SO}(8)$, ${\rm SO}(1,7)$ and ${\rm ISO}(7)$. \par We compute the mass spectra on these vacua, which turn out to be $\varphi$-independent, and determine the corresponding AdS-supermultiplet structure. Our analysis shows that, while there are several AdS ${\cal N}=8 \,\longrightarrow\, {\cal N}=3$ supersymmetry breaking patterns, only one, for each residual symmetry, seems to be dynamically realized in the full non-linear theory.\par As a last comment, vacua with residual ${\rm SO}(4)$ symmetry were investigated in \cite{Borghese:2013dja}. This analysis however missed the vacua discussed here since it restricted the ${\rm SO}(8)$ singlets to a sector which is invariant under a $D_4$ discrete subgroup of ${\rm SU}(8)$.\footnote{As a consequence, the gravitino mass matrix which is consistent with these symmetry requirements is proportional to the identity matrix and thus is different from the one we obtain for ${\cal N}=4$.}\par The paper is organized as follows. In Section \ref{sec:AdS} we formulate the problem of systematically studying the spontaneous $\mathcal{N}=8$ supersymmetry breaking on an AdS vacuum with residual extended supersymmetry: After a first introduction of the embedding tensor formalism, we consider the spontaneous supersymmetry breaking to $\mathcal{N}>2$ on AdS vacua and derive the corresponding system of quadratic equations on the non-vanishing components of the embedding tensor. We show that ${\cal N}>2$ residual supersymmetry requires the massive gravitinos to transform non-trivially under the associated ${\rm SO}({\cal N})$ R-symmetry group. In particular, we deduce the absence of solutions with $8>\mathcal{N}>4$ residual supersymmetry.\\ In section \ref{N83}, we study the possible AdS ${\cal N}=8 \,\longrightarrow\, {\cal N}=3$ supersymmetry breaking patterns at the level of the corresponding supersymmetry multiplets. In section \ref{sec:vacua}, we then describe the $\mathcal{N}=4$ and $\mathcal{N}=3$ classes of solutions to the quadratic equations. We identify, for the different values of the angular parameter $\varphi$, the corresponding gauge groups through the signature of the Cartan-Killing metric and by identifying the ${\rm E}_{7(7)}$-invariant quantities constructed out of the embedding tensor with the same quantities evaluated on $\omega$-rotated ${\rm SO}(8)$ \cite{DallAgata:2012bb}, ${\rm SO}(1,7)$ groups and on ${\rm ISO}(7)$. We show that these vacua disappear in the $\omega\rightarrow 0$ limit. Finally we give the AdS-supermultiplet structure and bosonic mass spectrum for the two classes of solutions. Appendix~\ref{app:not} summarizes our conventions and normalizations for the mass matrices; appendix~\ref{app:details} collects some of the computational details for the results of the main text. \section{AdS vacua with extended supersymmetry} \label{sec:AdS} \subsection{Gauged ${\cal N}=8$ supergravity} Let us briefly review some key formulas of gauged ${\cal N}=8$ supergravity, for details we refer to~\cite{deWit:2002vt,deWit:2007mt,LeDiffon:2011wt}. Gaugings of maximal ${\cal N}=8$ supergravity are described by the gauge group generators $X_{MN}{}^K$, ($M, N = 1, \dots, 56$) which in turn are obtained by contracting the ${\rm E}_{7(7)}$ generators $t_\alpha$ ($\alpha = 1, \dots, 133$) with a given embedding tensor $\Theta_M{}^\alpha$ \begin{eqnarray} X_{MN}{}^K &=& \Theta_M{}^\alpha\,(t_\alpha)_N{}^K \;. \label{XMNK} \end{eqnarray} They satisfy the quadratic identity \begin{eqnarray}{} [X_M, X_N] &=& -X_{MN}{}^K\,X_K \qquad\Longleftrightarrow\qquad \Omega^{MN} \Theta_M{}^\alpha \Theta_N{}^\beta ~=~ 0 \;, \label{Xalgebra} \end{eqnarray} which poses a quadratic constraint on the embedding tensor $\Theta_M{}^\alpha$, and exhibits the closure of the gauge algebra. The dressing of the generators (\ref{XMNK}) with the scalar dependent complex vielbein $\left\{{\cal V}_{M\,[ij]}, {\cal V}_M{}^{[ij]}\equiv({\cal V}_{M\,[ij]})^*\right\}$, $i, j =1, \dots, 8$, defines the $T$-tensor \begin{eqnarray} (T_{ij})^{klmn} &\equiv& \frac12\,({\cal V}^{-1})_{ij}{}^M ({\cal V}^{-1})^{kl}{}^N\,(X_{M})_{N}{}^K\,{\cal V}_K{}^{mn} \;,\qquad \mbox{etc.} \;. \label{TT} \end{eqnarray} The various components of this tensor will show up in the field equations of the gauged theory and parametrize the couplings. The fact that the embedding tensor $\Theta_M{}^\alpha$ is restricted to the ${\bf 912}$ representation of ${\rm E}_{7(7)}$ can be expressed by parametrizing the components of the $T$-tensor according to \begin{eqnarray} (T_{ij})_{kl}{}^{mn} &=& \frac12\left( \delta_{[k}{}^{[m} A{}^{n]}{}_{l]ij} + \delta_{i[k}{}^{mn} A_{l]j}-\delta_{j[k}{}^{mn} A_{l]i} \right) \;,\nonumber\\[.5ex] (T_{ij})^{rs}{}_{pq} &=& - \frac12\left(\delta_{[p}{}^{[r} A{}^{s]}{}_{q]ij} + \delta_{i[p}{}^{rs} A_{q]j}- \delta_{j[p}{}^{rs} A_{q]i}\right) \;,\nonumber\\[.5ex] (T_{ij})_{kl \, pq} &=& \frac{1}{24}\, \varepsilon_{klpqrstu} \, \delta_{[i}{}^r A_{j]}{}^{stu} \;,\nonumber\\[.5ex] (T_{ij})^{rs \; mn} &=& \delta_{[i}{}^{[r} A_{j]}{}^{smn]} \;, \label{TAB} \end{eqnarray} in terms of the scalar tensors\footnote{Here and in the following the coupling constant $g$ is absorbed in the definition of the tensors $A_{ij}$, $A_i{}^{jkl}$.} $A_{ij}$, $A_i{}^{jkl}$, satisfying $A_{[ij]}=0$, $A_{i}{}^{jkl} = A_{i}{}^{[jkl]}$, and $A_{i}{}^{jki}=0$. These tensors represent the ${\bf 36}$ and ${\bf 420}$ representations of ${\rm SU}(8)$, respectively, and parametrize the Yukawa-type couplings in the Lagrangian as \begin{eqnarray} {\cal L}_{\rm Yuk} &=& e\,\Big\{\frac12\sqrt{2}\, A_{1\,ij}\, \bar{\psi}^i_\mu \gamma^{\mu\nu} \psi^j_\nu + \frac{1}{6} A_{i}{}^{jkl} \, \bar{\psi}^i_\mu \gamma^\mu \chi_{jkl} \nonumber\\ &&{}\qquad\qquad + \frac1{144} \sqrt{2}\, \varepsilon^{ijkpqrlm}\,A{}^{n}{}_{pqr} \, \bar{\chi}_{ijk}\,\chi_{lmn} ~+~ {\rm h.c.}\Big \} \;, \label{Yuk} \end{eqnarray} for the eight gravitini $\psi_\mu{}^i$ and the 56 fermions $\chi_{ijk}$\,. The quadratic constraints (\ref{Xalgebra}) on the embedding tensor induce the following identities among the scalar dependent tensors $A_{ij}$, $A_i{}^{jkl}$ \begin{eqnarray} 0&=& A{}^k{}_{lij} \,A_{n}{}^{mij} - A_{l}{}^{kij} \, A{}^m{}_{nij} -4A{}^{(k}{}_{lni}A^{m)i}-4A_{(n}{}^{mki}A_{l)i} \nonumber\\ &&{}-2\,\delta_{l}^{m}\,A_{ni}A{}^{ki}+2\,\delta_{n}^{k}\,A_{li}A{}^{mi} \;,\nonumber \\[1ex] 0&=& A{}^i{}_{jk[m} \,A{}^k{}_{npq]} +A_{jk}\delta^{i}_{[m}A{}^{k}{}_{npq]} -A_{j[m}A^{i}{}_{npq]} \nonumber\\ && {}+\frac1{24}\,\varepsilon_{mnpqrstu}\, \left(A_{j}{}^{ikr}\, A_{k}{}^{stu} +A^{ik}\delta_{j}^{r}A_{k}{}^{stu}-A^{ir}A_{j}{}^{stu}\right) \;, \nonumber\\[1ex] 0&=& A^r{}_{ijk} \,A_{r}{}^{mnp} - 9\,A^{[m}{}_{r[ij}\, A_{k]}{}^{np]r} - 9\, \delta_{[i}{}^{[m} \, A^n{}_{|rs|j}\,A_{k]}{}^{p]rs} \nonumber\\ &&{}- 9\, \delta_{[ij}{}^{[mn}\,A^{|u|}{}_{k]rs}\, A_{u}{}^{p]rs} +\delta{}_{ijk}{}^{mnp} \,A^u{}_{rst} \, A_{u}{}^{rst} \;. \label{quad} \end{eqnarray} Let us finally note that the scalar potential of the theory is given in terms of these tensors by \begin{eqnarray} V&=& -\frac{3}{4}\,\left( A_{kl} A^{kl} - \frac{1}{18}\, A_{n}{}^{jkl} {A^n}_{jkl}\right) \;, \label{potential} \end{eqnarray} and that its extremal points are given by those values for the scalar fields at which the tensor \begin{eqnarray} {\cal C}_{ijkl}&=& {A{}^m}_{[ijk} A_{l]m} + \frac34\, {A{}^m}_{n[ij} {A^n}_{kl]m} \;, \label{defC} \end{eqnarray} becomes anti-selfdual: \begin{eqnarray} {\cal C}_{ijkl} + \frac1{24}\, \varepsilon_{ijklmnpq}\, {\cal C}^{mnpq} &=& 0\;. \label{ext} \end{eqnarray} At these extremal points, the couplings~(\ref{Yuk}) give rise to the fermionic mass terms. For example, the gravitino masses are obtained as the eigenvalues of the properly normalized tensor $A_{ij}$. For vanishing gauge fields and constant scalars, the Killing spinor equations of the theory reduce to \begin{eqnarray} 0 &\stackrel{!}{\equiv}& \delta_{\epsilon} \psi_{\mu}^i ~=~ 2\, {\cal D}_{\mu} \epsilon^i + \sqrt{2} \,A^{ij} \gamma_{\mu} \epsilon_j \;,\nonumber\\ 0 &\stackrel{!}{\equiv}&\delta_{\epsilon} \chi^{ijk} ~=~ -2 \,A_l{}^{ijk} \epsilon^l \;. \label{KSE} \end{eqnarray} Let us give, for the sake of completeness, the mass matrices for the various fields \cite{LeDiffon:2011wt}. The linearization of the scalar field equations yields, to lowest order, \begin{eqnarray} \Box\, \delta\phi_{ijkl} &=& {\cal M}_{ijkl}{}^{mnpq}\,\delta\phi_{mnpq} + {\cal O}(\delta\phi^2) \;, \end{eqnarray} where $\delta\phi_{ijkl}$ are fluctuations of the self-dual scalar fields $\phi_{ijkl}=\ft1{24}\varepsilon_{ijklpqrs}\, \phi^{pqrs}$ around their vacuum value $\phi_0=(\phi_0^{ijkl})$ and the scalar mass matrix ${\cal M}_{ijkl}{}^{mnpq}$ is given by \begin{eqnarray} {\cal M}_{ijkl}{}^{mnpq}\,\delta\phi^{ijkl}\delta\phi_{mnpq} &=& 6\, \left(A_{m}{}^{ijk}A^{l}_{ijn}\!-\!\ft14A_{i}{}^{jkl}A^{i}_{jmn}\right)\delta\phi^{mnpq}\delta\phi_{klpq} \nonumber\\ &&{} +\left(\ft5{24}\,A_{i}{}^{jkl}A^{i}{}_{jkl}-\ft12A_{ij}A^{ij}\right) \delta\phi^{mnpq}\delta\phi_{mnpq} \nonumber\\ &&{} -\ft23\,A_{i}{}^{jkl} A^{m}{}_{npq} \, \delta\phi^{inpq}\delta\phi_{jklm}\nonumber\\[1ex] &=& 12\,V^{(2)}(\delta\phi) \;, \label{Mscalar_sym} \end{eqnarray} where we have denoted by $V^{(2)}(\delta\phi)$ the terms on the scalar potential (\ref{potential}) which are second order in $\delta\phi$ upon expansion around the vacuum $\phi_0$: \begin{equation} V(\phi)~=~V_0+V^{(2)}(\delta\phi)+{\cal O}(\delta\phi^3)\,. \end{equation} The vector mass matrix reads \begin{eqnarray} {\cal M}_{\rm vec} &=& \left( \begin{array}{cc} {\cal M}_{ij}{}^{kl} & {\cal M}_{ijkl} \\ {\cal M}^{ijkl}& {\cal M}^{ij}{}_{kl} \end{array} \right) \;, \label{M_vector} \end{eqnarray} with \begin{eqnarray} {\cal M}_{ij}{}^{kl} &=& -\ft16\,A_{[i}{}^{npq} \delta_{j]}^{[k} A^{l]}{}_{npq} +\ft12\, A_{[i}{}^{pq[k}A^{l]}{}_{j]pq}\,, \nonumber\\[2ex] {\cal M}_{ijkl} &=& \ft1{36}\,A_{[i}{}^{pqr} \epsilon_{j]pqrmns[k} A_{l]}{}^{mns}\,. \end{eqnarray} We can also give this matrix a manifestly symplectic covariant form \begin{equation} {\cal M}_{{\rm vec}\,M}{}^N=\frac{1}{6}\,\left[{\rm Tr}(X_M\,X_P)+{\rm Tr}(\mathbb{M}^{-1}\,X_M\,\mathbb{M}\,(X_P)^T)\right]\,\mathbb{M}^{PN}\,, \end{equation} where $\mathbb{M}_{MN}=(\mathcal{V}\,\mathcal{V}^T)_{MN}$ is the symmetric, symplectic, positive definite matrix constructed from the coset representative $\mathcal{V}_M{}^N$ in the ${\bf 56}$ of ${\rm E}_{7(7)}$. By virtue of the quadratic constraint (\ref{Xalgebra}) on $\Theta$, the matrix ${\cal M}_{\rm vec}$ always has 28 vanishing eigenvalues (corresponding to the magnetic vector fields), while the remaining eigenvalues define the masses of the (electric) vector fields.\par Finally, the gravitino and fermion mass matrices are: \begin{eqnarray} {\cal M}_\psi{}^{ij} &=& \sqrt{2}\,A^{ij} \;,\qquad {\cal M}_\chi{}^{ijk,lmn} ~=~ \ft1{12} \sqrt{2}\, \epsilon^{ijkpqr[lm} A^{n]}{}_{pqr} \;. \label{Mferm} \end{eqnarray} The first matrix ${\cal M}_\psi$ carries the information about the breaking of supersymmetry and the latter matrix has to be evaluated after projecting out the fermions that are eaten by the massive gravitinos. Explicitly, at an AdS vacuum and in a basis in which $A_{ij}$ is diagonal, the effective fermion mass matrix is given by \begin{eqnarray} {\cal M}_\chi{}^{ijk,lmn} ~=~ \frac1{12} \sqrt{2}\, \Big( \epsilon^{ijkpqr[lm} A^{n]}{}_{pqr}~+~ \frac{4}{3}\, \sum_{p,q}{}^\prime A_p{}^{ijk} A_q{}^{lmn} \Big(\frac{A}{A^2 + {\bf 1}\, V/6}\Big)^{pq} \Big) \;, \end{eqnarray} with the sum running only over the massive gravitino directions. \subsection{${\cal N}>2$ AdS vacua} \label{subsec:N2} We have reviewed, how a given embedding tensor defines the scalar potential (\ref{potential}) of gauged supergravity which in turn may carry extremal points (\ref{ext}) at which supersymmetry is (partially) broken. The embedding tensor formalism allows to nicely invert the problem and to search for vacua with given properties by simultaneously scanning the set of all possible gaugings. That strategy has e.g.\ been applied in \cite{Dibitetto:2011gm,DallAgata:2011aa,Borghese:2012qm,Borghese:2012zs,Borghese:2013dja} in order to identify and analyze vacua with a given residual symmetry group. Concretely, any joint solution to the quadratic equations (\ref{quad}) and the vacuum condition (\ref{ext}) defines a vacuum in some maximal gauged supergravity. The associated embedding tensor and gauge group generators can then be restored via (\ref{TAB}), (\ref{TT}), and (\ref{XMNK}). In this section, we will investigate AdS vacua in maximal supergravity that preserve more than 2 supersymmetries. Let us assume that the matrices $A_{ij}$, $A_i{}^{jkl}$ describe an AdS vacuum preserving ${\cal N}$ supersymmetries, i.e.\ assume the existence of ${\cal N}$ independent solutions of (\ref{KSE}). Without loss of generality, we may then choose a basis $i=(\alpha, a)$ in which \begin{eqnarray} |A_{\alpha\beta}|&=&g\,\delta_{\alpha\beta}\;,\quad A_{\alpha a}~=~0 \;,\quad A{}^\alpha{}_{ijk}~=~0\;, \nonumber\\ &&{} \mbox{for}\quad \alpha,\beta = 1, \dots, {\cal N}\;,\quad a={\cal N}+1, \dots, 8 \;, \label{ansatz_susy} \end{eqnarray} and try to solve the quadratic equations (\ref{quad}), (\ref{ext}) under these assumptions for the remaining components of the tensors $A_{ij}$, $A_i{}^{jkl}$\,. First, let us note that for all non-vanishing ${\cal N}>0$, equations (\ref{ext}) follow directly as upon reduction of equations (\ref{quad}) by (\ref{ansatz_susy}). This is nothing but a remnant of the fact that the existence of a Killing spinor in general implies part of the remaining bosonic equations of motion (in this case the scalar field equations for constant scalars). It thus remains to solve equations (\ref{quad}) with the ansatz (\ref{ansatz_susy}). Since they are homogeneous, we may furthermore set $g=1$\,. Some contraction of the first equation from (\ref{quad}) then allows to deduce the value of the potential as \begin{eqnarray} V &=& -6\;. \end{eqnarray} On the other hand, the first equation of (\ref{quad}) with $k=\alpha, l=\beta, n=\gamma$ yields \begin{eqnarray} 0&=& -2A{}^{m}{}_{\beta \gamma \delta}\,A^{\alpha \delta} \quad \Longrightarrow\quad A{}^{m}{}_{\alpha\beta\gamma}~=~0 \;, \label{null3} \end{eqnarray} thus imposes the absence of the components $A{}^{m}{}_{\alpha\beta\gamma}$. For later use, we also note that the second equation of (\ref{quad}) in particular implies that \begin{eqnarray} 0 &=& 3\, A{}^a{}_{\alpha e[b} A{}^e{}_{cd]\beta} + A{}^a{}_{e\alpha\beta} A{}^e{}_{bcd}+ A_{\alpha\beta}\,A{}^a{}_{bcd}-\frac16\,\epsilon_{\alpha\beta \,bcd \, ijk} A^{ae} A_{e}{}^{ijk} \;. \label{q2l} \end{eqnarray} Let us now specialize to the case of ${\cal N}>2$ preserved supersymmetries. In this case, the preserved supercharges transform in the vector representation of the AdS R-symmetry ${\rm SO}({\cal N})$. We can then give a systematic discussion of these vacua according to the transformation of the broken supercharges (i.e.\ the massive gravitino fields) under that ${\rm SO}({\cal N})$. In particular, all non-vanishing components of the tensors $A_{ij}$, $A_i{}^{jkl}$ must be singlets under ${\rm SO}({\cal N})$. Let us consider as an example the case when all broken supercharges are singlet under ${\rm SO}({\cal N})$. If ${\cal N}>4$ this is the only option (in the absence of non-trivial ${\rm SO}({\cal N})$ representations of sufficiently small size). The non-vanishing components of the tensors $A_{ij}$, $A_i{}^{jkl}$ are thus given by \begin{eqnarray} \left\{ A_{\alpha\beta},\, A_{ab},\,A{}^a{}_{bcd} \right\} \;, \end{eqnarray} with all other possible singlets under ${\rm SO}({\cal N})$ vanishing, in view of (\ref{ansatz_susy}) and (\ref{null3}). Now (\ref{q2l}) immediately implies that also $A{}^a{}_{bcd}=0$, i.e.\ the entire tensor $A_i{}^{jkl}$ vanishes. Then however the first equation of (\ref{quad}) implies that \begin{eqnarray} A_{ac}\,A^{cb} &=& \delta_a{}^b \;, \end{eqnarray} i.e.\ after diagonalisation the eigenvalues of $A_{ab}$ are of absolute value 1 and all correspond to unbroken supersymmetries. The resulting vacuum thus is an ${\cal N}=8$ vacuum. We conclude that there are no ${\cal N}>2$ AdS vacua (other than the ${\cal N}=8$ ones) if the broken supercharges transform as singlets under the ${\rm SO}({\cal N})$ R-symmetry. In particular, there are no AdS vacua in maximal supergravity preserving $4<{\cal N}<8$ supersymmetries. For $\mathcal{N}=6$ AdS vacua, this is consistent with the result of \cite{Andrianopoli:2008ea}. In the following, we will thus assume that the broken supercharges transform non-trivially under the ${\rm SO}({\cal N})$ R-symmetry and determine the general solution of (\ref{quad}) for ${\cal N}>2$. \section{AdS ${\cal N}=8 \,\longrightarrow\, {\cal N}=3$ supersymmetry breaking patterns}\label{N83} \label{sec:N3patterns} \begin{table}[bt] \begin{center} {\small \begin{tabular}{|c||c|c|c|c|} \hline $\Delta$ $\Big\backslash$ $s$ & $\frac32$ & $1$ & $\frac12$ & $0$ \\ \hline\hline $E_0+3$ &&&&$[j]$ \\ \hline $E_0+\frac52$ &&&$[j+1]+[j]\,$\textcolor{blue}{$+\,[j-1]$}&\\ \hline $E_0+2$ && $[j+1]$\textcolor{blue}{$\,+\,[j]+[j-1]$}&& \begin{tabular}{c} $[j+2]+[j+1]+[j]$\\ \textcolor{blue}{${}+[j]+[j-1]+[j-2]$} \end{tabular} \\ \hline $E_0+\frac32$ &\textcolor{blue}{$[j]$}&& \begin{tabular}{c} $[j+2]+[j+1]$\textcolor{blue}{${}+[j+1]$}\\ \textcolor{blue}{${}+2[j]+2[j-1]+[j-2]$} \end{tabular}&\\ \hline $E_0+1$ &&\textcolor{blue}{$[j+1]+[j]+[j-1]$}&& \begin{tabular}{c} $[j+2]$\textcolor{blue}{$\,+\,[j+1]+2[j]$}\\ \textcolor{blue}{${}+[j-1]+[j-2]$} \end{tabular}\\ \hline $E_0+\frac12$ &&&\textcolor{blue}{$[j+1]+[j]+[j-1]$}&\\ \hline $E_0$ &&&&\textcolor{blue}{$[j]$}\\ \hline \end{tabular} } \caption{\small The long ${\cal N}=3$ gravitino multiplet $DS(3/2,E_0,j)_{\rm L}$, organized by energy~$\Delta$ and spin~$s$. When the energy saturates the unitarity bound $E_0=j+1$, the blue states in the table form a semi-short multiplet \textcolor{blue}{$DS(3/2,j+1,j)_{\rm S}$} and the other states decouple as a vector multiplet $DS(1,j+2)$.} \label{tab:Gmult} \end{center} \end{table} Before we start the analysis of the ${\cal N}=3$ solutions of the quadratic constraints (\ref{quad}), it is instructive to study the possible decompositions of the ${\cal N}=8$ supergravity multiplet into ${\cal N}=3$ multiplets, i.e.\ to identify the possible kinetic scenarios of supersymmetry breaking. The multiplet structure of the ${\cal N}=3$ AdS supergroup $OSp(3|4)$ is well known, see \cite{Freedman:1983na,Termonia:1999cs,Fre:1999xp}, in the following we will adopt the notation from \cite{Fre:1999xp}. The relevant multiplets for our discussion are the massive gravitino multiplets which accommodate the five massive gravitinos after the supersymmetry breaking ${\cal N}=8\longrightarrow {\cal N}=3$. The structure of the generic long gravitino multiplet $DS(3/2,E_0,j)_{\rm L}$ is recollected in table~\ref{tab:Gmult}. It is characterized by two numbers: the energy $E_0$ of its ground state, and the isospin $j$, characterising the representation of the gravitino under the R-symmetry group ${\rm SO}(3)$. Unitarity imposes the bound $E_0\ge j+1$ for the ground state energy (the ground state having spin 0). As usual for such supergroups, multiplet shortening occurs when the unitarity bound is saturated. At this value of $E_0$, the multiplet splits into a short massive gravitino multiplet together with a vector multiplet according to \begin{eqnarray} DS(3/2,E_0,j)_{\rm L}\Big|_{E_0 = j+1} &\longrightarrow& DS(3/2,j+1,j)_{\rm S} + DS(1,j+2) \;, \end{eqnarray} with the structure of the vector multiplet $DS(1,j+2)$ given in table~\ref{tab:Vmult}. At low values of $j$, the multiplet structure becomes non-generic, but the tables still capture the correct representation content upon formally extending the definition of ${\rm SO}(3)$ representations $[j]$ to negative $j$ according to\footnote{By $-[j-1]$ we mean that the isospin multiplet structure is obtained by deleting the representations with negative isospin ($[-j]$) and, for each of them, a representation $[j-1]$.} \begin{eqnarray} [-j] &\equiv& -[j-1]\;. \end{eqnarray} In particular, the lowest-lying short gravitino multiplet $DS(3/2,1,0)_{\rm S}$ carries a massless gravitino, three massless vectors, three fermions, and two scalars of energy $1$ and $2$. Due to the massless gauge fields, its presence in the spectrum implies an enhancement of supersymmetry and gauge symmetry. Similarly, the massless vector multiplet $DS(1,1)$ carries six scalars together with a massless vector and four fermions. \begin{table}[bt] \begin{center} {\small \begin{tabular}{|c||c|c|c|c|} \hline $\Delta$ $\Big\backslash$ $s$ & $1$ & $\frac12$ & $0$ \\ \hline\hline $j_0+2$ & && $[j_0-2]$ \\ \hline $j_0+\frac32$ && $[j_0-1]+[j_0-2]$ & \\ \hline $j_0+1$ &$[j_0-1]$ && $[j_0]+[j_0-1]+[j_0-2]$ \\ \hline $j_0+\frac12$ && $[j_0]+[j_0-1]$ &\\ \hline $j_0$ &&&$[j_0]$\\ \hline \end{tabular} } \caption{\small The ${\cal N}=3$ vector multiplet $DS(1,j_0)$.} \label{tab:Vmult} \end{center} \end{table} \begin{table}[bt] \begin{center} {\small \begin{tabular}{|c||c|c|c|c|} \hline $\Delta$ $\Big\backslash$ $s$ & $2^*$ & $\frac32^*$ & $1^*$ & $\frac12$ \\ \hline\hline $3$ &[0]&&& \\ \hline $\frac52$ &&[1]&& \\ \hline $2$ &&&[1]& \\ \hline $\frac32$ &&&&[0] \\ \hline \end{tabular} } \caption{\small The ${\cal N}=3$ massless gravity multiplet $DS(2,3/2,0)_{\rm S}$.} \label{tab:GN3} \end{center} \end{table} With the multiplet structure given in tables~\ref{tab:Gmult}, \ref{tab:Vmult}, the possible supersymmetry breaking patterns correspond to the different ways of splitting up the ${\cal N}=8$ supergravity multiplet into ${\cal N}=3$ multiplets. At this stage, we do not make any assumption about the energies of the various states (other than those implied by unitarity). The ${\cal N}=8$ supergravity multiplet consists of the graviton, 8 gravitinos, 28 vectors, 56 fermions and 70 scalars. Upon subtracting the ${\cal N}=3$ supergravity multiplet $DS(2,3/2,0)_{\rm S}$, given in table~\ref{tab:GN3}, we are left with 5 gravitinos, 25 vectors, 55 fermions and 70 scalars, to be packaged into ${\cal N}=3$ multiplets. There are various options for the splitting of the five massive gravitinos into ${\rm SO}(3)$ R-symmetry representations: \begin{eqnarray} {\rm I)} &5 ~\longrightarrow& 5\;,\nonumber\\ {\rm II)}&5 ~\longrightarrow& 3+1+1\;,\nonumber\\ {\rm III)}& 5 ~\longrightarrow& 2+2+1\;, \label{var5} \end{eqnarray} where we have taken into account the reality property of the gravitinos which rules out decompositions such as $4+1$, $3+2$, etc.. Moreover, the general discussion of section~\ref{subsec:N2} has ruled out the trivial decomposition $5\longrightarrow 1+1+1+1+1$\,. Let us discuss the patterns (\ref{var5}) one by one. Option I) in (\ref{var5}) leaves the five massive gravitinos in the irreducible spin-2 representation of ${\rm SO}(3)$. According to table~\ref{tab:Gmult}, they can sit either in a long multiplet $DS(3/2,E_0,2)_{\rm L}$ or in its shortened version $D(3/2,3,2)_{\rm S}$. Simple counting of states shows that the long multiplet carries 30 vector fields and thus does not fit into ${\cal N}=8$ supergravity. The short multiplet $DS(3/2,E_0,2)$ on the other hand does fit into ${\cal N}=8$ supergravity with the remaining states filling precisely two vector multiplets. A first possible kinetic pattern thus is given by \begin{eqnarray} {\rm I)} &:& {\cal N}=8\;\longrightarrow\; DS(2,3/2,0)_{\rm S} + DS(3/2,3,2)_{\rm S} + 2\cdot DS(1,1) \;. \label{optI} \end{eqnarray} The next option in (\ref{var5}) is the partition $3+1+1$, in which case there are various possibilities depending on the embedding of these gravitinos into the corresponding long or short gravitino multiplets. Naive counting allows for the following possibilities \begin{eqnarray} \begin{tabular}{c|cccc} II)&3&1&1&vectors\\\hline a):& L&S&S& 1\\ b):& S&L&L& 0 \\ c):& S&L&S& 3\\ d):& S&S&S& 6 \\ \end{tabular} \qquad\;. \label{optII} \end{eqnarray} The last column denotes the number of vector multiplets that describe the remaining matter spectrum, once the gravity and gravitino multiplets are subtracted from ${\cal N}=8$. Here, we note the following property of the vector multiplet: when ignoring the energy of the states, the field content of $DS(1,j_0)$ coincides with the tensor product of $DS(1,1)$ with the ${\rm SO}(3)$ representation $[j_0-1]$ of the vector fields. As a consequence, for instance the 3 vectors in the third row of (\ref{optII}) can either correspond to three multiplets $DS(1,1)$ or to a single multiplet $DS(1,2)$, the field content only differs in energies. Let us take a closer look at the decompositions of (\ref{optII}): the cases a) and d) both carry two gravitinos in the short massless $DS(3/2,1,0)_{\rm S}$, i.e.\ both cases in fact correspond to a supersymmetry enhancement to ${\cal N}=5$. Such vacua have been ruled out by the general discussion in section~\ref{sec:AdS} and cannot be dynamically realized. We are thus left with the options IIb) and IIc), of which the latter corresponds to a supersymmetry enhancement to ${\cal N}=4$. The third option in (\ref{var5}) is the partition $2+2+1$, for which we find two possibilities \begin{eqnarray} \begin{tabular}{c|cccc} III)&2&2&1&vectors\\\hline a):& S&S&L& 3\\ b):& S&S&S& 6 \\ \end{tabular} \qquad\;. \label{optIII} \end{eqnarray} In summary, the possible AdS ${\cal N}=8 \,\longrightarrow\, {\cal N}=3$ supersymmetry breaking patterns are given by the following decompositions of the ${\cal N}=8$ supergravity multiplet \begin{eqnarray} {\rm I)} &:& DS(2,3/2,0)_{\rm S} + DS(3/2,3,2)_{\rm S} + 2\cdot DS(1,1) \;, \label{opts} \\ {\rm IIb)} &:& DS(2,3/2,0)_{\rm S} + DS(3/2,2,1)_{\rm S} + DS(3/2,E_1,0)_{\rm L} + DS(3/2,E_2,0)_{\rm L} \;, \nonumber\\ {\rm IIc)} &:& DS(2,3/2,0)_{\rm S} + DS(3/2,2,1)_{\rm S}+ DS(3/2,E_0,0)_{\rm L}+ DS(3/2,1,0)_{\rm S} + DS(1,2) \;, \nonumber\\ {\rm IIIa)} &:& DS(2,3/2,0)_{\rm S} + 2\cdot DS(3/2,3/2,1/2)_{\rm S} + DS(3/2,E_0,0)_{\rm L} + DS(1,2) \;, \nonumber\\ {\rm IIIb)} &:& DS(2,3/2,0)_{\rm S} + 2\cdot DS(3/2,3/2,1/2)_{\rm S} + DS(3/2,1,0)_{\rm S} + 2\cdot DS(1,2) \;. \nonumber \end{eqnarray} In the following we will study which of these patterns can actually be dynamically realized in ${\cal N}=8$ supergravity and determine the specific gaugings which allow for the corresponding vacua. \section{${\cal N}=3$ and ${\cal N}=4$ vacua} \label{sec:vacua} \subsection{Solutions of the quadratic equations} \label{solutions} The ${\rm SO}(8)$ subgroup of ${\rm SU}(8)$ naturally splits into ${\rm SO}(3)\times {\rm SO}(5)$. We require the vacuum at the origin (and thus the tensors $A_{ij},\,A^i{}_{jkl}$) to be invariant under the diagonal group ${\rm SO}(3)_{\rm d}$ of the ${\rm SO}(3)$ group acting only on the Killing spinors and a second ${\rm SO}(3)$ embedded inside ${\rm SO}(5)$ according to the transformation of the massive gravitinos. We shall separately discuss the three cases corresponding to the allowed inequivalent embeddings (\ref{var5}) of ${\rm SO}(3)$ inside ${\rm SO}(5)$ and in each if them study the solutions to the system (\ref{quad}). In all cases we have reduced the system by implementing the most general ansatz in terms of singlets under ${\rm SO}(3)_{\rm d}$ and (with the help of mathematica) systematically scanned the remaining equations for their real solutions. Such solutions turn out to be extremely rare. Some computational details are relegated to appendix~\ref{app:details}. \subsubsection{Case ${\bf 5}\rightarrow {\bf 5}$} With this decomposition, there are six ${\rm SO}(3)_{\rm d}$ singlets in the tensors $A_{ij}$, $A_i{}^{jkl}$, three of which are killed by the general discussion of section~\ref{subsec:N2}. It is straightforward to verify that the remaining system of quadratic equations for three parameters does not possess any real solution (other than the known ${\cal N}=8$ solution), such that this possibility is ruled out by direct computation. \subsubsection{Case ${\bf 5}\rightarrow {\bf 2}+{\bf 2}+{\bf 1}$} Let us first split the $A,B,\dots$ indices into $\Lambda,\Sigma,\dots{}=4,5,6,7$ labeling the fundamental of the ${\rm SO}(4)$ inside ${\rm SO}(5)$, and identify the singlet in the decomposition with the value $i=8$. The index $i$ thus splits in $i=\alpha,\Lambda,8$. Next we embed ${\rm SO}(3)$ inside ${\rm SO}(4)$ by identifying its generators with the anti-self-dual matrices $(t^{(-)}_\alpha)^\Lambda{}_\Sigma$: \begin{align} t^{(-)}_1&=\left( \begin{array}{llll} 0 & \frac{1}{2} & 0 & 0 \\ -\frac{1}{2} & 0 & 0 & 0 \\ 0 & 0 & 0 & -\frac{1}{2} \\ 0 & 0 & \frac{1}{2} & 0 \end{array} \right)\,\,;\,\,\,t^{(-)}_2=\left( \begin{array}{llll} 0 & 0 & -\frac{1}{2} & 0 \\ 0 & 0 & 0 & -\frac{1}{2} \\ \frac{1}{2} & 0 & 0 & 0 \\ 0 & \frac{1}{2} & 0 & 0 \end{array} \right)\,\,;\,\,\,t^{(-)}_3=\left( \begin{array}{llll} 0 & 0 & 0 & -\frac{1}{2} \\ 0 & 0 & \frac{1}{2} & 0 \\ 0 & -\frac{1}{2} & 0 & 0 \\ \frac{1}{2} & 0 & 0 & 0 \end{array} \right) \end{align} We also have the complementary set of generators $({t}^{(+)}_{\hat{\alpha}})^\Lambda{}_\Sigma$, commuting with $t_\alpha$, obtained by changing the sign of the 4th row and columns of the latter. The following properties hold: \begin{equation} t^{(\pm)}_{\alpha \,\Lambda\Sigma}=\pm\frac{1}{2}\epsilon_{\Lambda\Sigma\Gamma\Delta}\,t^{(\pm)}_{\alpha \,\Gamma\Delta}\,. \end{equation} The ${\rm SO}(3)_{\rm d}$ generators in the ${\bf 8}$ of ${\rm SO}(8)$ read: \begin{equation} t_\alpha=\left(\begin{matrix}\epsilon_{\beta\alpha\gamma} & {\bf 0} & {\bf 0}\cr {\bf 0} & (t^{(-)}_{\alpha})_{\Lambda}{}^{\Sigma} & {\bf 0}\cr {\bf 0}& {\bf 0}& 0\end{matrix}\right)\,, \end{equation} and close the $\mathfrak{so}(3)_{\rm d}$ algebra: \begin{equation} [t_{\alpha},\,t_{\beta}]=\epsilon_{\alpha\beta\gamma}\,t_\gamma\,.\label{so3alg} \end{equation} With the general ansatz for $A_{ij}$, $A_i{}^{jkl}$ in terms of singlets under this ${\rm SO}(3)_{\rm d}$ (c.f.\ appendix~\ref{app:details}), we find aside from the known ${\cal N}=8$ solution $A_i{}^{jkl}=0,\,A_{77}=\pm 1,\,\,A_{88}=e^{i\varphi}$, only the following ${\cal N}=3$ solution: \begin{align} A_{\alpha\beta}&=\delta_{\alpha\beta}\;,\quad A_{\Lambda\Sigma}=\frac{3}{2}\,\epsilon\,\delta_{\Lambda\Sigma}\;,\quad A_{88}=-\sqrt{3}\,e^{3i\varphi}\;,\quad\nonumber\\ A^\Lambda{}_{\Sigma\alpha\beta}&=\epsilon_{\alpha\beta\gamma}\,(t^{(-)}_{\gamma})^\Lambda{}_{\Sigma}\;,\quad A^\Lambda{}_{\Sigma\alpha 8}=-\sqrt{3}\,e^{i\varphi}\,(t^{(-)}_{\alpha})^\Lambda{}_{\Sigma}\;,\quad,\nonumber\\ A^\Lambda{}_{\Sigma\Gamma\Delta}&=\epsilon \frac{\sqrt{3}}{2}\,e^{-i\varphi}\,\epsilon_{\Lambda\Sigma\Gamma\Delta}\;,\quad A^8{}_{\alpha\beta\gamma}=0\;,\quad A^8{}_{8\Lambda\Sigma}=0\;,\quad A^8{}_{\alpha\Lambda\Sigma}=-2\, \epsilon\,e^{-2i\varphi}\,\,t^{(-)}_{\alpha \,\Lambda\Sigma}\,,\label{s5221} \end{align} with real $\varphi$ and $\epsilon=\pm1$\,. It effectively depends only on the phase $\varphi$, since the sign $\epsilon=\pm 1 $ can be absorbed by an ${\rm SU}(8)$ transformation. \subsubsection{Case ${\bf 5}\rightarrow {\bf 3}+{\bf 1}+{\bf 1}$} Let us split the index $i$ into $i=\alpha,\alpha',a$, where $\alpha=1,2,3$, $\alpha'=4,5,6$ and $a=7,8$ is the index labeling the singlets. The ${\rm SO}(3)_{\rm d}$ generators in the ${\bf 8}$ of ${\rm SO}(8)$ read: \begin{equation} t_\alpha=\left(\begin{matrix}\epsilon_{\beta\alpha\gamma} & {\bf 0} & {\bf 0}\cr {\bf 0} & \epsilon_{\beta'\alpha\gamma'} & {\bf 0}\cr {\bf 0}& {\bf 0}& {\bf 0}_2\end{matrix}\right)\,, \end{equation} and satisfy the relations (\ref{so3alg}). With the general ansatz for $A_{ij}$, $A_i{}^{jkl}$ in terms of singlets under this ${\rm SO}(3)_{\rm d}$ (c.f.\ appendix~\ref{app:details}), we find aside from the known ${\cal N}=8$ solution only the following solution: \begin{align} A_{\alpha\beta}&=\delta_{\alpha\beta}\;,\quad A_{\alpha'\beta'}~=~2\xi\,\delta_{\alpha'\beta'} \;,\quad A_{77}=2\,\eta\;,\quad A_{88}=\xi\,\eta\,e^{i\varphi}\,,\nonumber\\ A^7{}_{\alpha'\beta'\gamma}&=\sqrt{2}\,\xi\,\eta\,e^{-i\frac{\varphi}{4}}\, \epsilon_{\alpha'-3\,\beta'-3\,\gamma}\;,\quad A^{\alpha'}{}_{\beta'\gamma 7}=-\sqrt{2}\,e^{-i\frac{\varphi}{4}}\,\epsilon_{\alpha'-3\,\beta'-3\,\gamma}\,,\nonumber\\ A^{\alpha'}{}_{\beta'\gamma' 8}&=\sqrt{2}\,\xi\,e^{i\frac{\varphi}{4}}\,\epsilon_{\alpha'-3\,\beta'-3\,\gamma'-3}\;,\quad A^{\alpha'}{}_{\alpha ab}=-\eta\,e^{i\frac{\varphi}{2}}\,\delta^{\alpha'-3}_\alpha\,\epsilon_{ab}\,,\nonumber\\ A^7{}_{8\alpha'\alpha}&=-\xi\,e^{i\frac{\varphi}{2}}\,\delta_{\alpha'-3\,\alpha}\;,\quad A^{\alpha'}{}_{\beta' \alpha\beta}=-2\,\delta^{\alpha'-3\,\beta'-3 }_{\alpha\beta}\,,\label{s53111} \end{align} where $\eta,\,\xi=\pm 1$. The parameter $\xi$ can be disposed of by means of a ${\rm SU}(8)$ transformation while the sign $\eta$ can be changed by shifting $\varphi\rightarrow \varphi+2\pi$. We can thus set $\xi=\eta=+1$. Notice that $A^8{}_{ijk}=0$ which implies that this is actually an ${\cal N}=4$ solution and that the residual symmetry group is enhanced to ${\rm SO}(4)$. \subsection{Gauge Groups and ${\rm E}_{7(7)}$-Invariants} We have identified two AdS vacua in maximal supergravity by solving the system of quadratic constraints (\ref{quad}) for the embedding tensor. As the next step, we will have to determine the associated gauge groups, i.e.\ identify in which gauged maximal supergravity these vacua live. We can compute the associated gauge group generators via (\ref{TAB}), (\ref{TT}), and (\ref{XMNK}). Much of the structure of the gauge group can already be inferred from the ${\rm E}_{7(7)}$-invariant signature of the (generalized) Cartan-Killing metric \begin{equation} \mbox{sign}[{\rm Tr}(X_M\cdot X_N)]\,. \label{CK} \end{equation} The above matrix has $28$ vanishing eigenvalues (due to the locality constraint (\ref{Xalgebra})) while the other $28$ eigenvalues define the Cartan-Killing metric of the gauge algebra. \subsubsection{The ${\cal N}=4$ vacuum} We first compute the Cartan-Killing metric (\ref{CK}) for the ${\cal N}=4$ vacuum (\ref{s53111}) as a function of the angular parameter $\varphi$. This allows the following identification of the corresponding underlying gauge group: \begin{eqnarray} \begin{tabular}{|c|c|c|} \hline parameter & signature of C.-K. metric & gauge group\\\hline $\varphi=2\pi $& $(1_+,15_-,12_0)$ & $[{\rm SO}(1,1)\times {\rm SO}(6)]\ltimes T^{12}$ \\ $0\le \varphi< 2\pi$&$(7_+,21_-)$ & ${\rm SO}(1,7)$ \\ \hline \end{tabular} \label{tab4} \end{eqnarray} where $T^{12}$ denotes a subgroup generated by twelve nilpotent operators. Notice that AdS vacua in theories with gauged ${\rm SO}(1,7)$ and $[{\rm SO}(1,1)\times {\rm SO}(6)]\ltimes T^{12}$ groups were found in \cite{DallAgata:2011aa}. The residual supersymmetry, symmetry group (${\rm SO}(4)$) and spectrum of our vacuum distinguishes it from those found in the same reference. The gauge group ${\rm SO}(1,7)$ alone is not sufficient to determine the gauged supergravity, since there is a one-parameter class of such theories~\cite{DallAgata:2012bb,DallAgata:2014ita}. Rather, we expect to find a mapping between the angular parameter $\varphi$ which defines our vacua, and the $\omega$-angle that labels the one-parameter class of ${\rm SO}(1,7)$ theories. To this end, we compute other ${\rm E}_{7(7)}$-invariants on the ${\cal N}=4$ vacuum, to compare with the same quantities evaluated for the $\omega$-rotated ${\rm SO}(1,7)$ gauge group. In particular, we consider the $1540\times 1540$ matrix: \begin{equation} \mathbb{K}_{MN}{}^{PQ}\equiv \frac{1}{4}\,d^{R_1 R_2 R_3 R_4}\,X_{R_1 M}{}^K\,X_{R_2NK}\,X_{R_3}{}^{PL}\,X_{R_4 L}{}^Q\,, \label{K} \end{equation} quartic in the gauge group generators (\ref{XMNK}), which is antisymmetric in $[MN]$ and $[PQ]$, by virtue of the total symmetry of the ${\rm E}_{7(7)}$-invariant tensor $d^{R_1 R_2 R_3 R_4}$. This tensor $\mathbb{K}$ is related to one computed in \cite{DallAgata:2012bb} for the $\omega$-deformed ${\rm SO}(8)$ gauging. Instead of evaluating the eigenvalues of this matrix, as was done for the corresponding tensor in \cite{DallAgata:2012bb} for a specific gauging, we evaluate the traces of its powers. We compute them for our ${\cal N}=4$ solution and for the ($\omega$-deformed) ${\rm SO}(1,7)$ and ${\rm SO}(8)$ gauging. For the invariant $d$-tensor we use the following form in the ${\rm SU}(8)$-basis: \begin{align} d^{MNPQ}\,\Lambda_M\Lambda_N\Lambda_P \Lambda_Q&=\Lambda^{i_1 i_2}\Lambda^{i_3 i_4}\Lambda^{i_5 i_6}\Lambda^{i_7 i_8}\,\epsilon_{i_1\dots i_8}+\Lambda_{i_1 i_2}\Lambda_{i_3 i_4}\Lambda_{i_5 i_6}\Lambda_{i_7 i_8}\,\epsilon^{i_1\dots i_8}+\nonumber\\ &+96\,{\rm Tr}(\Lambda\bar{\Lambda}\Lambda\bar{\Lambda})-24\,{\rm Tr}(\Lambda\bar{\Lambda})^2\,,\nonumber\\ \Lambda_M& \equiv(\Lambda,\bar{\Lambda}) ~\equiv~ (\Lambda_{ij},\Lambda^{ij})\;. \end{align} For the ${\rm SO}(8)$ and ${\rm SO}(1,7)$ gaugings the $X$-tensor (\ref{XMNK}) is computed via (\ref{TT}), (\ref{TAB}), starting from fermion shift tensors of the form \cite{DallAgata:2011aa}: \begin{equation} A_{ij}=e^{i\,\omega}\,{\rm Tr}(\theta)\,\delta_{ij}\,\,\,,\,\,\,\,A_{i}{}^{jkl}=e^{-i\,\omega}\,(\Gamma_i{}^{jkl})_{IJ}\theta^{IJ}\,, \end{equation} where \begin{equation} \theta^{IJ}={\rm diag}(1,1,1,1,1,1,1,\kappa)\,, \end{equation} with $\kappa=+1$ for ${\rm SO}(8)$ and $-1$ for ${\rm SO}(1,7)$. We find for the traces of the various powers of (\ref{K}) \begin{align} {\rm Tr}(\mathbb{K})&=0\,,\nonumber\\ {\rm Tr}(\mathbb{K}^2)&=2^{23}\times 3^4\times 5\times 7\times\left(7 (5 \kappa +3)+28 (\kappa -1) \cos (4 \omega )+(\kappa +7) \cos (8 \omega )\right)\,,\nonumber\\ {\rm Tr}(\mathbb{K}^3)&=2^{36}\times 3^7\times 5\times 7\times \left(35 \kappa +4 (7 \kappa +1) \cos (4 \omega )+(\kappa -1) \cos (8 \omega )-3\right) \sin ^2(2 \omega )\,,\nonumber\\ {\rm Tr}(\mathbb{K}^4)&\propto {\rm Tr}(\mathbb{K}^2)^2\,.\label{trKn} \end{align} Notice that in the ${\rm SO}(8)$ case these invariants have half-period $\pi/8$, namely they assume all possible values in the interval $\omega\in(0,\pi/8)$, while for the ${\rm SO}(1,7)$ gauging the half-period is $\pi/4$. In the former case we have independent gaugings only for $\omega\in(0,\pi/8)$, while in the latter case for $\omega\in(0,\pi/4)$, consistently with the results of \cite{DallAgata:2012bb,DallAgata:2012sx}. Eq.s (\ref{trKn}) do not hold for $\kappa=0$, corresponding to ${\rm ISO}(7)$, in which case all traces are zero.\par On our ${\cal N}=4$ solution (\ref{s53111}), (denoting the corresponding tensor by $\mathbb{K}_s$) these traces become \begin{align} {\rm Tr}(\mathbb{K}_s)&=0\,,\nonumber\\ {\rm Tr}(\mathbb{K}_s^2)&=2^{6}\times 3^4\times 5\times 7\times \left(3 \cos \left(\frac{\varphi }{2}\right)+\cos (\varphi )+2\right)\,,\nonumber\\ {\rm Tr}(\mathbb{K}_s^3)&=-2^{10}\times 3^7\times 5\times 7\times \cos ^4\left(\frac{\varphi }{4}\right)\,\nonumber\\ {\rm Tr}(\mathbb{K}_s^4)&\propto {\rm Tr}(\mathbb{K}_s^2)^2\,. \label{tr4} \end{align} These expressions are symmetric under $\varphi\rightarrow -\varphi$ and $\varphi\rightarrow \varphi+4\pi$, so that they assume all possible values in the interval $(0,2\pi)$. Following the same reasoning as for the ${\rm SO}(1,7)$ and ${\rm SO}(8)$ cases, we can then argue that $X$ tensors in this class with generic $\varphi$ can be ${\rm SU}(8)$-rotated to one with $\varphi\in (0,2\pi)$. In comparing the traces for the ${\cal N}=4$ vacuum (\ref{tr4}) to those of the ${\rm SO}(1,7)$ gauging (\ref{trKn}), we assume that \begin{equation} X^{(s)}_{MN}{}^P=\lambda(\varphi)\,{\rm E}_{7(7)}\star(X_{MN}{}^P)\,,\label{XsX} \end{equation} where $X^{(s)}_{MN}{}^P$ is the $X$-tensor on the ${\cal N}=4$ vacuum, ${\rm E}_{7(7)}\star(X_{MN}{}^P)$ is the ${\rm E}_{7(7)}$-rotated $X$ tensor of the ${\rm SO}(1,7)$ gauging, and we allowed for a proportionality factor $\lambda(\varphi)$ depending on the parameter $\phi$. Clearly the traces of $\mathbb{K}$ do not depend on the ${\rm E}_{7(7)}$-rotation, so that eq. (\ref{XsX}) implies: \begin{align} {\rm Tr}(\mathbb{K}_s^2)&=\lambda(\varphi)^8\,{\rm Tr}(\mathbb{K}^2)\,\,\,\,,\,\,\,\,\,{\rm Tr}(\mathbb{K}_s^3)=\lambda(\varphi)^{12}\,{\rm Tr}(\mathbb{K}^3)\,.\label{KsK} \end{align} We immediately realize that for $\omega=0$ the above system has no solution since ${\rm Tr}(\mathbb{K}_s^2)=0$ while ${\rm Tr}(\mathbb{K}^2)\neq 0$, thus implying $\lambda=0$. Similarly for $\varphi=2\pi$, the second of (\ref{KsK}) implies $\omega=0$ while in the first ${\rm Tr}(\mathbb{K}_s^2)=0$ while ${\rm Tr}(\mathbb{K}^2)\neq 0$, againg implying $\lambda=0$. This is compatible with our finding (\ref{tab4}) that the gauge group at $\varphi=2\pi$ degenerates to $[{\rm SO}(1,1)\times {\rm SO}(6)]\ltimes T^{12}$. For generic values of $\omega$ and $\varphi$ the relation between the two parameters can then be deduced from the ($\lambda$-independent) equation \begin{equation} \frac{{\rm Tr}(\mathbb{K}_s^3)^2}{{\rm Tr}(\mathbb{K}_s^2)^3}=\frac{{\rm Tr}(\mathbb{K}^3)^2}{{\rm Tr}(\mathbb{K}^2)^3}\,\,\Longleftrightarrow\,\,\,\,\frac{9 \cos ^2\left(\frac{\varphi }{4}\right)}{70 \left(2 \cos \left(\frac{\varphi }{2}\right)+1\right)^3}=\frac{144\, (\cos (4 \omega )+3)^4 \sin ^4(2 \omega )}{35\, (-28 \cos (4 \omega )+3 \cos (8 \omega )-7)^3}\;. \label{num4} \end{equation} whose solution $\omega(\varphi)$ is plotted in Fig.~\ref{N4}. We conclude that the ${\cal N}=4$ vacuum can be found (for generic values of $\varphi$), only in the $\omega$-rotated ${\rm SO}(1,7)$ gauging, while in the limit $\omega\rightarrow 0$ it disappears. At the corresponding point ${\varphi=2\pi}$ in the parameter space of solutions it turns into a vacuum of the gauging with non-semisimple gauge group $[{\rm SO}(1,1)\times {\rm SO}(6)]\ltimes T^{12}$. \begin{figure}[ht!] \centering \includegraphics[scale=.6]{omegalambdaN4.pdf} \caption{{\small The parameters $\omega$ (blue) and $\lambda$ (red) as function of $\varphi$ for the ${\cal N}=4$ vacuum. The gauge group is ${\rm SO}(1,7)$ except for the point $\varphi=2\pi$ where $\lambda$ vanishes and the gauge group degenerates to $[{\rm SO}(1,1)\times {\rm SO}(6)]\ltimes T^{12}$. }} \label{N4} \end{figure} \subsubsection{The ${\cal N}=3$ vacuum} The same analysis can be repeated for the ${\cal N}=3$ vacuum (\ref{s5221}). In this case the computation of the Cartan-Killing metric (\ref{CK}) for the gauge group indicates the following correspondence between the values of $\varphi$ and the gauge group \begin{eqnarray} \begin{tabular}{|c|c|c|} \hline parameter & signature of C.-K. metric & gauge group\\\hline $0\le \varphi<\frac{\pi}{6} $& $(0_+,28_-)$ & ${\rm SO}(8)$ \\ $\frac{\pi}{6} < \varphi\le\pi $&$(7_+,21_-)$ & ${\rm SO}(1,7)$ \\ $ \varphi=\frac{\pi}{6}$&$(0_+,21_-, 7_0)$ & ${\rm ISO}(7)$ \\ \hline \end{tabular} \label{tab3} \end{eqnarray} Computing the traces of the tensor (\ref{K}) on our ${\cal N}=3$ solution we find in this case: \begin{align} {\rm Tr}(\mathbb{K}_s)&=0\,,\nonumber\\ {\rm Tr}(\mathbb{K}_s^2)&=-2^{-1}\times 3^8\times 5\times 7\times \cos (\varphi ) \left(\sqrt{3} (\cos (2 \varphi )+3)-7 \cos (\varphi )\right)\,,\nonumber\\ {\rm Tr}(\mathbb{K}_s^3)&=2^{-2}\times 3^{11}\times 5\times 7\times (24 \sqrt{3} \cos (\varphi )-18 \cos (2 \varphi )+2 \sqrt{3} \cos (3 \varphi )-27)\,,\nonumber\\ {\rm Tr}(\mathbb{K}_s^4)&\propto {\rm Tr}(\mathbb{K}_s^2)^2\,, \label{tr3} \end{align} which should now be compared to (\ref{trKn}) for $\kappa=\pm1$ in the different intervals of (\ref{tab3}). The expressions of (\ref{tr3}) are symmetric under $\varphi\rightarrow -\varphi$ and $\varphi\rightarrow \varphi+2\pi$, so that they assume all possible values in the interval $(0,\pi)$. We then argue that an $X$-tensor in this class with a generic $\varphi$ can be ${\rm SU}(8)$-rotated to one within $\varphi\in (0,\pi)$. The correspondence between $\varphi$ and $\omega$ for the $\omega$-rotated ${\rm SO}(1,7)$ and ${\rm SO}(8)$ groups is obtained from an equation analogous to (\ref{num4}) and plotted in Fig. \ref{N3}. This illustrates that $\lambda$ vanishes, as $\omega\rightarrow 0$ ($\varphi\rightarrow \pi/6$), so that the vacuum disappears from the corresponding gauged theories. At this point in the $\varphi$ parameter space it becomes a vacuum of an ${\rm ISO}(7)$ gauged theory. Indeed, for $\varphi=\pi/6$, all traces of (\ref{tr3}) vanish.\\ \begin{figure}[ht!] \centering \includegraphics[scale=.6]{omegalambda78.pdf} \caption{{\small The parameters $\omega$ (blue) and $\lambda$ (red) as function of $\varphi$ for the ${\cal N}=3$ vacuum. The gauge group is ${\rm SO}(8)$ for $\varphi<\pi/6$ and ${\rm SO}(1,7)$ for $\varphi>\pi/6$. A the point $\varphi=\pi/6$ where $\lambda$ vanishes, the gauge group degenerates to ${\rm ISO}(7)$. }} \label{N3} \end{figure} The existence of the proportionality parameter $\lambda(\varphi)$ depending on $\varphi$ (or, equivalently on $\omega$) is due to the fact that we have fixed the value the potential in our vacua to a given value ($-6$) by choosing the coupling constant. This function therefore encodes the dependence, in the corresponding $\omega$-rotated theories, of the cosmological constant of these vacua on $\omega$, which is a generic feature of all extrema of the potential aside from the $\mathcal{N}=8$ one \cite{DallAgata:2012bb,Borghese:2012zs}. \subsection{Mass spectra} We can eventually compute the spectra around the new vacua by evaluating the mass formulas (\ref{Mscalar_sym}), (\ref{M_vector}), and (\ref{Mferm}) for our solutions $A_{ij}$, $A_i{}^{jkl}$ and compare the result to the general multiplet structure discussed in section~\ref{sec:N3patterns}. We find that in all cases the spectra are independent of the parameter $\varphi$. \subsubsection{The ${\cal N}=3$ vacuum} The scalar mass spectrum on the ${\cal N}=3$ vacuum is: \begin{align} m^2\,L_0^2\qquad&: 1\times\left(3(1+\sqrt{3})\right)\,;\,\,6\times\left(1+\sqrt{3}\right)\,;\,\,1\times\left(3(1-\sqrt{3})\right)\,;\,\, 6\times\left(1-\sqrt{3}\right)\,;\nonumber\\ &4\times\left(-\frac{9}{4}\right)\,;\,\,18\times\left(-2\right)\,;\,\,12\times\left(-\frac{5}{4}\right)\,;\,\,22\times\left(0\right)\,, \end{align} in units of the inverse anti- de Sitter radius $1/L_0$ from (\ref{L0}). The Breitenlohner-Freedman bound $m^2\,L_0^2\ge -\frac{9}{4}$ \cite{Breitenlohner:1982jf} is satisfied by virtue of supersymmetry. The normalized vector masses are given by: \begin{align} m^2\,L_0^2\quad&: 3\times\left(3+\sqrt{3}\right)\,;\,\,3\times\left(3-\sqrt{3}\right)\,;\,\,4\times\left(\frac{15}{4}\right)\,;\,\, 12\times\left(\frac{3}{4}\right)\,;\,\,6\times\left(0\right)\,, \end{align} The 22 massless scalar fields are the Goldstone bosons for the massive vector fields. Together, we conclude that the ${\cal N}=3$ vacuum realizes option IIIa) from (\ref{opts}) with one long spin $3/2$ multiplet of energy $E_0=\sqrt{3}$ \begin{eqnarray} DS(2,3/2,0)_{\rm S} + 2\cdot DS(3/2,3/2,1/2)_{\rm S} + DS(3/2,\sqrt{3},0)_{\rm L} + 3\cdot DS(1,1) \end{eqnarray} The three extra massless vectors describe an extra ${\rm SO}(3)$ symmetry. Explicit computation of the fermionic mass matrices (\ref{Mferm}) also confirms the multiplet structure. \subsubsection{The ${\cal N}=4$ vacuum} The scalar mass spectrum on the ${\cal N}=4$ vacuum is: \begin{align} m^2\,L_0^2&:1\times\left(10\right)\,;\,\,10\times\left(4\right)\,;\,\,11\times\left(-2\right)\,;\,\,48\times\left(0\right)\,. \end{align} The vector masses are \begin{align} m^2\,L_0^2&: 7\times\left(6\right)\,;\,\,15\times\left(2\right)\,;\,\,6\times\left(0\right)\,, \end{align} 22 of the massless scalar fields are the Goldstone bosons for the massive vector fields, while the six massless vectors gauge the residual ${\rm SO}(4)$ group. This solution thus realizes option IIc) from (\ref{opts}) with a long spin $3/2$ multiplet of energy $E_0=2$ \begin{eqnarray} DS(2,3/2,0)_{\rm S}+ DS(3/2,1,0)_{\rm S} + DS(3/2,2,1)_{\rm S}+ DS(3/2,2,0)_{\rm L} + DS(1,2) \;, \end{eqnarray} and supersymmetry enhancement to ${\cal N}=4$, under which the first two multiplets combine into the ${\cal N}=4$ massless supergravity multiplet and the remaining three multiplets combine into a single ${\cal N}=4$ massive spin $3/2$ multiplet. Again, an explicit computation of the fermionic mass matrices (\ref{Mferm}) confirms this multiplet structure. \section{Conclusions} In this paper we have studied AdS vacua of maximal supergravity in four dimensions with residual $\mathcal{N}>2$ supersymmetry. We exclude on general grounds $8>\mathcal{N}>4$ vacua and find two 1-parameter classes of $\mathcal{N}=3$ and $4$ vacua, which can be embedded only in the $\omega$-rotated gauged models. Of particular importance are the models with ${\rm SO}(8)$ gauging since they exhibit in addition an $\mathcal{N}=8$ vacuum. The eleven dimensional origin of the latter is still debated and in \cite{DallAgata:2012bb} it was conjectured to corrrespond to certain to ABJ theories \cite{Aharony:2008gk}, through the AdS/CFT duality \cite{Maldacena:1997re}. Understanding the higher dimensional origin of the new $\mathcal{N}=3$ and $4$ vacua is an important problem which deserves investigation.\par Still in the light of the AdS/CFT correspondence, these new vacua should describe conformal fixed points of some dual (three-dimensional) field theory. It would be also interesting, in this respect, to study RG flows between the conformal critical points dual to the two kinds of vacua in the $\omega$-deformed ${\rm SO}(1,7)$ models, or interpolating between the $\mathcal{N}=8$ and $\mathcal{N}=3$ vacua in the $\omega$-deformed ${\rm SO}(8)$ theories, thus generalizing the analysis of \cite{Guarino:2013gsa,Tarrio:2013qga}.\par An other issue which deserves investigation is the study of black holes asymptoting the new $\mathcal{N}=3$ and $4$ vacua, along the lines of \cite{Anabalon:2013eaa}. It would also be interesting to understand to which extend the methods developed in this paper can be extended to a systematic analysis of the AdS (and Minkowski) vacuum in maximal supergravity with ${\cal N}=2$ and ${\cal N}=1$ supersymmetry. \section*{Acknowledgements} We wish to thank D. Roest for helpful and inspiring discussions. \bigskip \bigskip \bigskip
\part{Use this type of header for very long papers only} \maketitle \section{Introduction} \label{intro} \noindent Dynamical systems is concerned with the study of the asymptotic behavior of the orbits of a given system. Certain hypothesis like Smale's hyperbolicity guarantee the knowledge of this behavior. Indeed, the celebrated {\em Smale spectral decomposition theorem} asserts that every hyperbolic system on a compact manifold comes equipped with finite many pairwise disjoint compact invariant sets (homoclinic classes or singularities) to which every trajectory converge. Although present in a number of interesting examples, such a hypothesis is far from being abundant in the dynamical forrest. This triggered several attempts to extend it including the {\em sectional-hyperbolicity} \cite{memo}, committed to merge the hyperbolic theory to the so-called {\em geometric} and {\em multidimensional Lorenz attractors} \cite{abs}, \cite{bpv}, \cite{gw}. A number of results from the hyperbolic theory have been carried out to the sectional-hyperbolic context. This is nowadays matter of study in a number of works, see \cite{ap} and references therein. One of these results was motivated by the well-known fact that two different homoclinic classes contained in a common hyperbolic set are disjoint. It was quite natural to ask if this statement is also true in the sectional-hyperbolic context too. In other words, are two different homoclinic classes contained in a common sectional-hyperbolic set disjoint? But recent results dealing with this question say that the answer is negative \cite{mp2}, \cite{mp3}. Moreover, \cite{mp3} studied the dynamics of nontransitive sectional-Anosov flows with dense periodic orbits nowadays called {\em venice masks}. It was proved that three-dimensional venice masks with a unique singularity exists \cite{bmp} and that their maximal invariant set consists of two different homoclinic classes with nonempty intersection \cite{mp3}. Venice mask with $n$ singularities can be constructed for $n\geq3$ whereas ones with just two singularities have not been constructed yet. These fruitful results motivate a related problem which is the analysis of the intersection of a sectional-hyperbolic set for the flow and a sectional-hyperbolic set for the reversed flow. For simplicity we keep the terms {\em positively} and {\em negatively sectional-hyperbolic} for these sets (respectively) which was coined by Shy, Gan and Wen in their recent paper \cite{sgw}. After observing that every hyperbolic set can be realized as such an intersection, we show an example where such an intersection is not hyperbolic. Next we show that such an intersection is hyperbolic if the intersecting sets are both transitive. In general the intersection consists of a nonsingular hyperbolic set (possibly empty), finitely many singularities and regular orbits joining them. Finally, we construct a three-dimensional star flow exhibiting two homoclinic classes, one being positively (but not negatively) sectional-hyperbolic and the other being negatively (but not positively) sectional-hyperbolic, whose intersection reduces to a single periodic orbit. This will provide a counterexample to a conjecture by Zhu, Gan and Wen \cite{zgw} (as amended by Shy, Gan and Wen \cite{sgw}). \section{Statement of the results} \noindent Let $M$ be a differentiable manifold endowed with a Riemannian metric $\langle\cdot,\cdot\rangle$ an induced norm $\|\cdot\|$. We call {\em flow} any $C^1$ vector field $X$ with induced flow $X_t$ of $M$. If $\dim(M)=3$, then we say that $X$ is a {\em three-dimensional flow}. We denote by $\operatorname{Sing}(X)$ the set of singularities (i.e. zeroes) of $X$. By a {\em periodic point} we mean a point $x\in M$ for which there is a minimal $t>0$ satisfying $X_t(x)=x$. By an {\em orbit} we mean $O(x)=\{X_t(x):t\in\mathbb{R}\}$ and by a {\em periodic orbit} we mean the orbit of a periodic point. We say that $\Lambda\subset M$ is {\em invariant} if $X_t(\Lambda)=\Lambda$ for all $t\in\mathbb{R}$. In such a case we write $\Lambda^*=\Lambda\setminus \operatorname{Sing}(X)$. We say that $\Lambda\subset M$ is {\em transitive} if there is $x\in \Lambda$ such that $\omega(x)=\Lambda$, where $\omega(x)$ is the {\em $\omega$-limit set}, $$ \omega(x)=\left\{ y\in M:y=\lim_{n\to\infty} X_{t_n}(x)\mbox{ for some sequence }t_n\to\infty\right\}. $$ The {\em $\alpha$-limit set} $\alpha(x)$ is the $\omega$-limit set for the reversed flow $-X$. If the set of periodic points of $X$ in $\Lambda$ is dense in $\Lambda$, we say that $\Lambda$ has {\em dense periodic points}. A compact invariant set $\Lambda$ is {\em hyperbolic} if there is a continuous invariant splitting $T_\Lambda M=E^s\oplus E^X\oplus E^u$ and positive numbers $K,\lambda$ such that \begin{enumerate} \item $E^s$ is {\em contracting}, i.e., $\| DX_t(x)v^s_x\|\leq Ke^{-\lambda t}\|v^s_x\|$ for every $x\in\Lambda$, $v^s_x\in E^s_x$ and $t\geq0$. \item $E^X_ x$ is the subspace generated by $X(x)$ in $T_x M$, for every $x\in \Lambda$. \item $E^u$ is {\em expanding}, i.e., $\| DX_t(x)v^u_x\|\geq K^{-1}e^{\lambda t}\|v^u_x\|$ for every $x\in\Lambda$, $v^u_x\in E^u_x$ and $t\geq0$. \end{enumerate} A singularity or periodic orbit is hyperbolic if it does as a compact invariant set of $X$. The elements of a (resp. hyperbolic) periodic orbit will be called (resp. hyperbolic) periodic points. A singularity or periodic orbit is a {\em sink} (resp. {\em source}) if its unstable subbundle $E^u$ (resp. stable subbundle $E^s$) vanishes. Otherwise we call it {\em saddle type}. The invariant manifold theory \cite{hps} asserts that through any point $x$ of a hyperbolic set it passes a pair of invariant manifolds, the so-called stable and unstable manifolds $W^{s}(x)$ and $W^{u}(x)$, tangent at $x$ to the subbundles $E^s_x$ and $E^u_x$ respectively. Saturating them with the flow we obtain the weak stable and unstable manifolds $W^{ws}(x)$ and $W^{wu}(x)$ respectively. On the other hand, a compact invariant set $\Lambda$ {\em has a dominated splitting with respect to the tangent flow} if there are an invariant splitting $T_\Lambda M=E\oplus F$ and positive numbers $K,\lambda$ such that $$ \| DX_t(x)e_x\|\cdot \|f_x\| \leq Ke^{-\lambda t} \| DX_t(x)f_x\|\cdot\|e_x\|, \quad\quad\forall x\in \Lambda, t\geq0, (e_x,f_x)\in E_x\times F_x. $$ Notice that this definition allows every compact invariant set $\Lambda$ to have a dominated splitting with respect to the tangent flow: Just take $E_x=T_xM$ and $F_x=0$ for every $x\in \Lambda$ (or $E_x=0$ and $F_x=T_xM$ for every $x\in\Lambda$). However, such splittings need not to exist under certain constraints. For instance, not every compact invariant set has a dominated splitting $T_\Lambda M=E\oplus F$ with respect to the tangent flow which is {\em nontrivial}, i.e., satisfying $E_x\neq0\neq F_x$ for every $x\in \Lambda$. A compact invariant set $\Lambda$ is {\em partially hyperbolic} if it has a {\em partially hyperbolic splitting}, i.e., a dominated splitting $T_\Lambda M=E\oplus F$ with respect to the tangent flow whose dominated subbundle $E$ is contracting in the sense of (1) above. The Riemannian metric $\langle\cdot,\cdot\rangle$ of $M$ induces a {\em $2$-Riemannian metric} \cite{mv}, $$ \langle u,v/w\rangle_p=\langle u,v\rangle_p\cdot \langle w,w\rangle_p-\langle u,w\rangle_p\cdot \langle v,w\rangle_p, \quad\forall p\in M,\forall u,v,w\in T_pM. $$ This in turns induces a {\em $2$-norm} \cite{g} (or {\em areal metric} \cite{ka}) defined by $$ \|u,v\|=\sqrt{\langle u,u/v\rangle_p}, \,\,\,\,\,\,\forall p\in M,\forall u,v\in T_pM. $$ Geometrically, $\|u,v\|$ represents the area of the paralellogram generated by $u$ and $v$ in $T_pM$. If a compact invariant set $\Lambda$ has a dominated splitting $T_\Lambda M=E\oplus F$ with respect to the tangent flow, then we say that its central subbundle $F$ is {\em sectionally expanding} (resp. {\em sectionally contracting}) if $$ \| DX_t(x) u, DX_t(x) v\|\geq K^{-1}e^{\lambda t}\| u,v\|, \quad\quad\forall x\in \Lambda, u, v\in F_x, t\geq0. $$ (resp. $$ \| DX_t(x) u, DX_t(x) v\|\leq Ke^{-\lambda t}\| u,v\|, \quad\quad\forall x\in \Lambda, u, v\in F_x, t\geq0.) $$ By a {\em sectional-hyperbolic splitting} for $X$ over $\Lambda$ we mean a partially hyperbolic splitting $T_\Lambda M=E\oplus F$ whose central subbundle $F$ is sectionally expanding. Now we define sectional-hyperbolic set. \begin{defi} A compact invariant set $\Lambda$ is {\em sectional-hyperbolic} for $X$ if its singularities are hyperbolic and if there is a sectional-hyperbolic splitting for $X$ over $\Lambda$. Following \cite{sgw} we use the term {\em positively} (resp. {\em negatively) sectional-hyperbolic} to indicate a sectional-hyperbolic set for $X$ (resp. $-X$). The corresponding sectional-hyperbolic splitting will be termed {\em positively} (resp. {\em negatively}) {\em sectional-hyperbolic splitting}. \end{defi} This definition is slightly different from the original one given in Definition 2.3 of \cite{memo} (which requires, for instance, that the central subnbundle be two-dimensional at least). Such a difference permits {\em every} hyperbolic set $\Lambda$ to be both positively and negatively sectional-hyperbolic. Indeed, if $T_\Lambda M=E^s\oplus E^X\oplus E^u$ is the respective hyperbolic splitting, then $T_\Lambda M=E^s\oplus E^{se}$ with $E^{se}=E^X\oplus E^u$ and $T_\Lambda M=\hat{E}^s\oplus \hat{E}^{se}$ with $\hat{E}^s=E^u$ and $\hat{E}^{se}=E^s\oplus E^X$ define positively and negatively sectional-hyperbolic splittings respectively over $\Lambda$. In particular, {\em every hyperbolic set is the intersection of a positively and a negatively sectional-hyperbolic set}. \medskip One can ask if the hyperbolic sets are the sole possible intersection between a positively and a negatively sectional-hyperbolic set, but they aren't. In fact, there are nonhyperbolic compact invariant sets which, nevertheless, are both positively and negatively sectional-hyperbolic. This is the case of the example described in Figure \ref{fig1}. In such a figure $O(x)$ represents the orbit of $x\in W^s(\sigma_1)\cap W^u(\sigma_2)$ whereas a singularity of a three-dimensional flow is {\em Lorenz-like} for $X$ if it has three real eigenvalues $\lambda_1,\lambda_2,\lambda_3$ satisfying $\lambda_2<\lambda_3<0<-\lambda_3<\lambda_1$. \begin{figure}[htv] \begin{center} \input{vicosa01.pdf_t} \caption{Nonhyperbolic but positively and negatively sectional-hyperbolic.}\label{fig1} \end{center} \end{figure} This counterexample motivates the search of sufficient conditions under which the intersection of a positively and a negatively sectional-hyperbolic set be hyperbolic. Our first result is about this problem. \begin{theorem} \label{th2} The intersection of a transitive positively sectional-hyperbolic set and a transitive negatively sectional-hyperbolic set is hyperbolic. \end{theorem} Consequently, \begin{corollary} \label{th1} Every transitive set which is both positively and negatively sectional-hyperbolic is hyperbolic. \end{corollary} The similar results replacing transitivity by denseness of periodic orbits hold. By looking at Figure \ref{fig1} we observe that this example consists of two singularities and a regular point $x$ whose $\omega$-limit and $\alpha$-limit set is a singularity. This observation is the motivation for the result below. \begin{theorem} \label{th3} Every compact invariant set which is both positively and negatively sectional-hyperbolic is the disjoint union of a (possibly empty) nonsingular hyperbolic set $H$, a (possibly empty) finite set of singularities $S$ and a (possibly empty) set of regular points $R$ such that $\alpha(x)\subset H\cup S$ and $\omega(x)\subset H\cup S$ for every $x\in R$. \end{theorem} Since the intersection of a positively and a negatively sectional-hyperbolic set is both positively and negatively sectional-hyperbolic, we obtain the following corollary. \begin{corollary} \label{c3} The intersection of a positively and a negatively sectional-hyperbolic set is a disjoint union of a (possibly empty) nonsingular hyperbolic set $H$, a (possibly empty) finite set of singularities $S$ and a (possibly empty) set of regular points $R$ such that $\alpha(x)\subset H\cup S$ and $\omega(x)\subset H\cup S$ for every $x\in R$. \end{corollary} Our next result is an example of nontrivial transitive sets which are positively and negatively sectional-hyperbolic (resp.) whose intersection is the simplest possible, i.e., a single periodic orbit. Denote by $\operatorname{Cl}(\cdot)$ the closure operation. We say that $H\subset M$ is a {\em homoclinic class} if there is a hyperbolic periodic point $x$ of saddle type such that $$ H=\operatorname{Cl}(\{q\in W^{ws}(x)\cap W^{wu}(x): \dim(T_qW^{ws}(x)\cap T_qW^{wu}(x))=1\}). $$ It follows from the {\em Birkhoff-Smale Theorem} that every homoclinic class is a transitive set with dense periodic orbits. Given points $x,y\in M$, if for every $\epsilon>0$ there are sequences of points $\{x_i\}_{i=0}^n$ and times $\{t_i\}_{i=0}^{n-1}$ such that $x_0=x$, $x_n=y$, $t_i\geq1$ and $d(X_{t_i}(x_i),x_{i+1})<\epsilon$ for every $0\leq i\leq n-1$, then we say that $x$ {\em is in the chain stable set of $y$}. If $x$ is in the chain stable set of $y$ and viceversa, then one says that $x$ and $y$ are {\em chain related}. If $x$ is chain related to itself, one says that $x$ is a {\em chain recurrent point}. The set of chain recurrent points is the {\em chain recurrent set} denoted by $CR(X)$. It is clear that the chain related relation is in equivalence on $CR(X)$. By using this equivalence, one splits $CR(X)$ into equivalence classes denominated {\em chain recurrent classes}. A flow is {\em star} if it exhibits a neighborhood $\mathcal{U}$ (in the space of $C^1$ flows) such that every periodic orbit or singularity of every flow in $\mathcal{U}$ is hyperbolic. With these definitions we obtain the following result. \begin{theorem} \label{thB} There is a star flow $X$ in the sphere $S^3$ whose chain recurrent set is the disjoint union of two periodic orbits $O_1$ (a sink), $O_2$ (a source); two singularities $s_-$ (a source), $s_+$ (a saddle); and two homoclinic classes $H_-$, $H_+$ with the following properties: \begin{itemize} \item $H_-$ is negatively (but not positively) sectional-hyperbolic; \item $H_+$ is positively (but not negatively) sectional-hyperbolic; \item $H_-\cap H_+$ is a periodic orbit. \end{itemize} \end{theorem} Recall that the {\em nonwandering set} of a flow $X$ is defined as the set of points $x\in M$ such that for every neighborhood $U$ of $x$ and $T>0$ there is $t\geq T$ satisfying $X_t(U)\cap U\neq\emptyset$. Given a certain subset $O$ of the space of $C^1$ flows, we say that a $C^1$ generic flow in $O$ satisfies another property (Q) if there is a residual subset of flows $R$ of $O$ such that every flow in $R$ satisfying (P) also satisfies (Q). \medskip There are two current conjectures relating star flows and sectional-hyperbolicity. These are based on previous results in the literature e.g. \cite{gaw}, \cite{mp}. \begin{conjecture}[Zhu-Shy-Gan-Wen \cite{sgw},\cite{zgw}] \label{zgw} The chain recurrent set of {\em every} star flow is the {\em disjoint} union of a positively sectional-hyperbolic set and a negatively sectional-hyperbolic set. \end{conjecture} \begin{conjecture}[Arbieto \cite{am}] The nonwandering set of a {\em $C^1$ generic} star flow is the disjoint union of finitely many transitive sets which are positively or negatively sectional-hyperbolic. \end{conjecture} However, the union $H_-\cup H_+$ of the homoclinic classes $H_-$ and $H_+$ in Theorem \ref{thB} is a chain recurrent class of the corresponding flow $X$ (because $H_-\cap H_+\neq\emptyset$). Therefore, Theorem \ref{thB} gives a counterexample for Conjecture \ref{zgw} in dimension $3$. Similar counterexamples can be obtained in dimension $\geq3$. \begin{corollary} \label{c2} There is a star flow in $S^3$ whose chain recurrent set is not the disjoint union of a positively sectional-hyperbolic set and a negatively sectional-hyperbolic set. \end{corollary} Another interesting feature regarding this counterexample is the existence of a chain recurrent class without any nontrivial dominated splitting with respect to the tangent flow. Moreover, every ergodic measure supported on this class is hyperbolic saddle. These features are related to \cite{cs} or \cite{m}. Notice also that the star flow in Corollary \ref{c2} can be $C^1$ approximated by ones exhibiting the heteroclinic cycle obtained by joinning the unstable manifold $W^u(\sigma_1)$ of $\sigma_1$ to the stable manifold $W^s(\sigma_2)$ of $\sigma_2$ in Figure \ref{fig1}. Such a cycle was emphasized in the figure after the statement of Lemma 3.3 in p.951 of \cite{zgw}. This put in evidence the role of robust transitivity in the proof of such a lemma. \section*{Acknowledgments} \noindent This paper grew out of discussions between authors and participants of the course {\em Topics in Dynamical Systems II} given at the Federal University of Rio de Janeiro, Brazil, in the last half of 2014. The authors express their gratitute to these participants including professors J. Aponte, T. Catalan, A.M. Lopez B. and H. Sanchez. The results in this paper were announced in the {\em I Workshop on Sectional-Anosov flows} which took place in September 22 of 2014 at the Federal University of Vi\c cosa-MG, Brasil. The authors would like to thank professors E. Apaza and B. Mejia for the corresponding invitation. \section{Proof of theorems \ref{th2} and \ref{th3}} \noindent First we prove Theorem \ref{th3}. For this we use the following technical definition. \begin{defi} A compact invariant set $\Lambda$ of a flow $X$ is {\em almost hyperbolic} if: \begin{enumerate} \item Every singularity in $\Lambda$ is hyperbolic. \item There are continuous invariant subbundles $E^s, E^u$ of $T_\Lambda M$ such that $E^s$ is contracting, $E^u$ is expanding and $$ T_{\Lambda^*}M=E^s\oplus E^X\oplus E^u. $$ \end{enumerate} \end{defi} Notice that this definition is symmetric with respect to the reversing-flow operation. Moreover, hyperbolic sets are almost hyperbolic but not conversely by the example in Figure \ref{fig1}. Likewise sectional-hyperbolic sets, the almost hyperbolic sets satisfy \begin{lemma}[Hyperbolic Lemma] Every compact invariant subset without singularities of an almost periodic set is hyperbolic. \end{lemma} More properties will be obtained from the lemma below. We denote by $B(x,\delta)$ the open $\delta$-ball operation, $\delta>0$. If $\sigma\in Sing(X)$ is hyperbolic, then we denote by $W^s_\delta(\sigma)$ (resp. $W^u_\delta(\sigma)$) the connected component of $B(\sigma,\delta)\cap W^s(\sigma)$ (resp. $B(\sigma,\delta)\cap W^u(\sigma)$) containing $\sigma$. \begin{lemma} \label{thA'} For every almost hyperbolic set $\Lambda$ of a flow $X$ there is $\delta>0$ such that $\Lambda\cap B(\sigma,\delta)\subset W^u_\delta(\sigma)\cup W^u_\delta(\sigma)$ for every $\sigma\in \operatorname{Sing}(X)\cap \Lambda$. \end{lemma} \begin{proof} It suffices to prove that if $x_n\in \Lambda^*$ is a sequence converging to some singularity $\sigma\in \Lambda$, then $x_n\in W^s(\sigma)\cup W^u(\sigma)$ for $n$ large enoch. Let $T_\sigma M=F^s_\sigma\oplus F^u_\sigma$ be the hyperbolic splitting of $\sigma$. By definition $T_{x_n}M=E^s_{x_n}\oplus E_{x_n}^X\oplus E^u_{x_n}$ so $$ \dim(E^s_{x_n})+\dim(E^u_{x_n})=\dim(M)-1, \quad\quad\forall n. $$ Passing to the limit we obtain $$ \dim(E^s_{\sigma})+\dim(E^u_{\sigma})=\dim(M)-1. $$ Since $E^s_\sigma$ and $E^u_\sigma$ are contracting and expanding respectively, we obtain $E^s_\sigma\subset F^s_\sigma$ and $E^u_\sigma\subset F^u_\sigma$. If $\dim(F^s_\sigma)>\dim(E^s_\sigma)+1$ we would have $$ \dim(E^u)=\dim(M)-1-\dim(E^s_\sigma)>\dim(M)-\dim(F^s_\sigma)=\dim(F^u_\sigma), $$ which is impossible. Then $\dim(F^s_\sigma)\leq \dim(E^s_\sigma)+1$. Analogously, $\dim(F^u_\sigma)\leq \dim(E^u_\sigma)+1$. Therefore, $\dim(F^s_\sigma)=\dim(E^s_\sigma)$ or $\dim(E^s_\sigma)+1$. Analogously $\dim(F^u_\sigma)=\dim(E^u_\sigma)$ or $\dim(E^u_\sigma)+1$. But we cannot have $\dim(F^s_\sigma)=\dim(E^s_\sigma)+1$ and $\dim(F^u_\sigma)=\dim(E^u_\sigma)+1$ simultaneously because $$ \dim(M)=\dim(F^s_\sigma)+\dim(F^u_\sigma) =\dim(E^s_\sigma)+\dim(E^u_\sigma)+2=\dim(M)+1 $$ which is absurd. All together imply $$ \dim(E^s_\sigma)=\dim(F^s_\sigma)\quad \quad\mbox{ or } \quad\quad \dim(E^u_\sigma)=\dim(F^u_\sigma). $$ Suppose $\dim(E^s_\sigma)=\dim(F^s_\sigma)$. If $y\in \Lambda\cap(W^s(\sigma)\setminus\{\sigma\})$ is sufficiently close to $\sigma$, then $\dim(E^s_y)=\dim(E^s_\sigma)=\dim(F^s_\sigma)=\dim(T_yW^s(\sigma))$. On the other hand, $E^s$ is contracting thus $E^s_y\subset T_yW^s(\sigma)$. From these remarks we obtain that if $\dim(E^s_\sigma)=\dim(F^s_\sigma)$, then $E^s_y=T_yW^s(\sigma)$ for all $y\in \Lambda\cap( W^s(\sigma)\setminus\{\sigma\})$ close to $\sigma$. Analogously if $\dim(E^u_\sigma)=\dim(F^u_\sigma)$, then $E^u_y=T_yW^u(\sigma)$ for all $y\in \Lambda\cap(W^u(\sigma)\setminus\{\sigma\})$ close to $\sigma$. Now suppose by contradiction that $x_n\not\in W^s(\sigma)\cup W^u(\sigma)$ for all $n$ (say). Then, by flowing the orbit of $x_n$ nearby $\sigma$, as described in Figure \ref{fig2}, we obtain two sequences $x_n^s,x^u_n$ in the orbit of $x_n$ such that $x^s_n\to y^s$ and $x^u_n\to y^u$ for some $y^s\in W^s(\sigma)\setminus \{\sigma\}$ and $y^u\in W^u(\sigma)\setminus\{\sigma\}$ close to $\sigma$. \begin{figure}[htv] \begin{center} \input{vicosa2-5-11.pdf_t} \caption{Proof of Lemma \ref{thA'}}\label{fig2} \end{center} \end{figure} If $\dim(E^s_\sigma)=\dim(F^s_\sigma)$ then $E^s_{y^s}=T_{y^s}W^s(\sigma)$ but also $E^X_{y^s}\subset T_{y^s}W^s(\sigma)$ since $W^s(\sigma)$ is an invariant manifold. Therefore, $E^X_{y^s}\subset E^s_{y^s}$ and then $E^X_{y^s}=0$ since the sum $T_{y^s}M=E^s_{y^s}\oplus E^X_{y^s}\oplus E^u_{y^s}$ is direct. This is a contradiction. Analogously we obtain a contradiction if $\dim(E^u_\sigma)=\dim(F^u_\sigma)$ and the proof follows. \end{proof} Now we relate sectional and almost hyperbolicity. \begin{lemma} \label{thA} Every compact invariant set which is both positively and negatively sectional-hyperbolic is almost hyperbolic. \end{lemma} \begin{proof} Let $\Lambda$ be a compact invariant set which is both positively and negatively sectional-hyperbolic. Then, every singularity in $\Lambda$ is hyperbolic. Moreover, there are positively and negatively sectional-hyperbolic splittings $$ T_\Lambda M=E^s\oplus E^{se}, \quad\mbox{ and }\quad T_\Lambda M=\hat{E}^s\oplus \hat{E}^{se}, $$ Taking $E^u=\hat{E}^s$ and $E^{sc}=\hat{E}^{se}$ we obtain an expanding and a sectional contracting subbundles of $T_\Lambda M$. Since $E^s$ is contracting, we have $E^X\subset E^{se}$ by Lemma 3.2 in \cite{as}. Similarly, $E^X\subset E^{sc}$ so $$ E^X\subset E^{se}\cap E^{sc}. $$ On the other hand, since $E^s$ is contracting and $E^u$ expanding, the angle $\langle E^s, E^u\rangle$ is bounded away from zero. Then, the dominating condition implies $$ E^u\subset E^{se} \quad\quad\mbox{ and } \quad\quad E^s\subset E^{sc}. $$ From this we have $T_\Lambda M=E^{se}+ E^{sc}$ and so $$ \dim(M)=\dim(E^{se})+\dim(E^{sc})-\dim(E^{se}\cap E^{sc}). $$ At regular points we cannot have a vector outside $E^X$ contained in $E^{se}\cap E^{sc}$. Then, $$ E^X=E^{se}\cap E^{sc} \quad\quad\mbox{ and so }\quad\quad \dim(E^{se}\cap E^{sc})=1 $$ in $\Lambda^*$. Replacing above we get $$ \dim(M)=\dim(E^{se})+\dim(E^{sc})-1. $$ But we also have $\dim(M)=\dim(E^u)+\dim(E^{sc})$ so $$ \dim(E^u)=\dim(E^{se})-1. $$ Since $E^u$ is expanding, we have $E^u\cap E^X=\{0\}$ thus $$ T_{\Lambda^*} M=E^s\oplus E^X\oplus E^u $$ proving the result. \end{proof} \begin{proof}[Proof of Theorem \ref{th1}] Let $\Lambda$ be the intersection of a positively and a negatively sectional-hyperbolic set of a flow $X$. Then, it is both positively and negatively sectional-hyperbolic and so almost hyperbolic by Lemma \ref{thA}. From this we can select $\delta>0$ as in Lemma \ref{thA'}. Clearly we can take $\delta$ such that the balls $B(\sigma,\delta)$ are pairwise disjoint for $\sigma\in S$, where $S=\operatorname{Sing}(X)\cap \Lambda$. Define $$ H=\displaystyle\bigcap_{(t,\sigma)\in\mathbb{R}\times S}X_t(\Lambda\setminus B(\sigma,\delta)) $$ and $R=\Lambda\setminus (H\cup S)$. Clearly $S$ consists of finitely many singularities. Moreover, $H$ is nonsingular hence hyperbolic by the Hyperbolic Lemma. Now take $x\in R$. Then, there is $(t,\sigma)\in\mathbb{R}\times S$ such that $X_t(x)\in B(\sigma,\delta)$. By Lemma \ref{thA'} we obtain $X_t(x)\in W^s(\sigma)\cup W^u(\sigma)$ hence $x\in W^s(\sigma)\cup W^u(\sigma)$. If $x\in W^s(\sigma)$ we obtain $\omega(x)\subset H\cup S$. If $X_r(x)\notin \cup_{\sigma\in S}B(\sigma,\delta)$ for all $r\leq 0$ then $\alpha(x)\subset H$. Otherwise, there is $(r,\rho)\in \mathbb{R}\times S$ such that $X_r(x)\in B(\rho,\delta)$ and so $x\in W^u(\rho)$. All together yields $\alpha(x)\subset H\cup S$. Similarly we have $\alpha(x)\subset H\cup S$ and $\omega(x)\subset H\cup S$ if $x\in W^u(\sigma)$ and the result follows. \end{proof} To prove Theorem \ref{th2} we use the following lemma. Recall that an invariant set is {\em nontrivial} if it does not reduces to a single orbit. \begin{lemma} \label{palilla} Let $\Lambda$ be a nontrivial transitive positively sectional-hyperbolic set of a flow $X$. If $\sigma\in\operatorname{Sing}(X)\cap \Lambda$, then the hyperbolic and the respective hyperbolic and positively sectional-hyperbolic splittings $T_\sigma M=F^s_\sigma\oplus F^u_\sigma$ and $T_\sigma M=E^s_\sigma\oplus E^{se}_\sigma$ of $\sigma$ satisfy $\dim(E^{se}_\sigma\cap F^s_\sigma)=1$. \end{lemma} \begin{proof} Clearly $E^s_\sigma\subset F^s_\sigma$. Suppose for a while that $E^s_\sigma=F^s_\sigma$. Then, $\dim(E^s_y)=\dim(T_yW^s(\sigma))$ for every $y\in \Lambda\cap W^s(\sigma)$ close to $\sigma$. As clearly $E^s_y\subset T_yW^s(\sigma)$ for all such points $y$, we obtain $E^s_y=T_yW^s(\sigma)$ for every $y\in \Lambda\cap W^s(\sigma)$ close to $\sigma$. On the other hand, we also have that $E^X_y\subset T_y W^s(\sigma)$ for all such points $y$. From this we conclude that $E^X_y\subset E^s_y$ for every point $y\in\Lambda\cap W^s(\sigma)$ close to $\sigma$. Now we observe that since $\Lambda$ is transitive we obtain $E^X\subset E^{se}$. Using again that $\Lambda$ is nontrivial transitive (see Figure \ref{fig2}) we obtain $y=y^s\in \Lambda^*\cap W^s(\sigma)$ close to $\sigma$. For such a point we obtain $0\neq E^X_y\subset E^s_y\cap E^{se}_y$ which is absurd. Therefore, $E^s_\sigma\neq F^s_\sigma$. Next we observe that $\dim(E^{se}_\sigma\cap F^s_\sigma)\leq1$ by sectional expansivity. Suppose for a while that $\dim(E^{se}_\sigma\cap F^s_\sigma)=0$. Clearly $E^s_\sigma\cap F^u_\sigma=0$ and so $F^u_\sigma\subset E^{se}_\sigma$ by domination. From this we obtain $T_\sigma M=E^{se}_\sigma\oplus F^s_\sigma$ thus $\dim(E^{se}_\sigma)+\dim(F^s_\sigma)=\dim(M)=\dim(F^s_\sigma)+\dim(F^u_\sigma)$ yielding $\dim(E^{se}_\sigma)=\dim(F^u_\sigma)$ so $E^{se}_\sigma=F^u_\sigma$ thus $E^s_\sigma=F^s_\sigma$ which is absurd. Therefore, $\dim(E^{se}_\sigma\cap F^s_\sigma)=1$ and we are done. \end{proof} \begin{proof}[Proof of Theorem \ref{th2}] Let $\Lambda_+$ and $\Lambda_-$ be transitive sets of a flow $X$ such that $\Lambda_+$ is positively sectional hyperbolic and $\Lambda_-$ is negatively sectional-hyperbolic. If one of these sets reduces to a single orbit, then the intersection $\Lambda_-\cap \Lambda$ reduces to that orbit and the result follows. So, we can assume both $\Lambda_+$ and $\Lambda_-$ are nontrivial. Let $T_{\Lambda_+}M=E^s\oplus E^{se}$ and $T_{\Lambda_-}M=\hat{E}^s\oplus \hat{E}^{se}$ be the positively and negatively sectional-hyperbolic splittings of $\Lambda_+$ and $\Lambda_-$ respectively. Denoting $E^u=\hat{E}^s$ and $E^{sc}=\hat{E}^{se}$ we obtain an expanding subbundle and a sectionally contracting subbundle of $T_\Lambda M$. Suppose for a while that there is $\sigma\in \Lambda_-\cap\Lambda_+\cap\operatorname{Sing}(X)$. By Lemma \ref{palilla} applied to $X$, we have that $\sigma$ has a real negative eigenvalues $\lambda^s$ corresponding to the one-dimensional eigendirection $E^{se}_\sigma\cap F^s_\sigma$. Similarly, applying the lemma to $-X$, we obtain a real positive eigenvalue $\lambda^u$ corresponding to the one-dimensional eigendirection $E^{sc}_\sigma\cap F^u_\sigma$. Take unitary vectors $v^s\in E^{se}_\sigma\cap F^s_\sigma$ and $v^u\in E^{sc}_\sigma\cap F^u_\sigma$. Since $$ (E^{se}_\sigma\cap F^s_\sigma)\cap (E^{sc}_\sigma\cap F^u_\sigma)\subset F^s_\sigma\cap F^u_\sigma=0, $$ we have that $v^s$ and $v^u$ are linearly independent. Then, $\|v^s,v^u\|\neq0$. Since $F^u_\sigma\subset E^{se}_\sigma$, we have $v^s,v^u\in E^{se}_\sigma$ so $$ e^{\lambda^st}e^{\lambda^ut}\|v^s,v^u\|=\|DX_t(\sigma)v^s,DX_t(\sigma)v^u\|\to\infty \quad\mbox{ as }\quad t\to\infty $$ by sectionally expansiveness. Then $$ \lambda^s+\lambda^u>0. $$ Similarly, since $F^s_\sigma\subset E^{sc}_\sigma$, we have $v^s,v^u\in E^{sc}_\sigma$ so $$ e^{-\lambda^st}e^{-\lambda^ut}\|v^s,v^u\|=\|DX_{-t}(\sigma)v^s,DX_{-t}(\sigma)v^u\|\to\infty \quad\mbox{ as }\quad t\to\infty $$ by sectionally expansiveness with respect to $-X$. Then, $$ \lambda^s+\lambda^u<0 $$ which is absurd. We conclude that $\Lambda_-\cap\Lambda_+\cap\operatorname{Sing}(X)=\emptyset$. Now we can apply the hyperbolic lemma for sectional-hyperbolic sets to obtain that $\Lambda_-\cap\Lambda_+$ is hyperpolic. This finishes the proof. \end{proof} \section{Proof of Theorem \ref{thB}} \noindent Roughly speaking, the proof consists of glueing the so-called {\em singular horseshoe} \cite{lp} with its time reversed counterpart. \begin{figure}[h] \begin{center} \input{vicosa110.pdf_t} \caption{}\label{fig3} \end{center} \end{figure} We star with the standard {\em Smale horseshoe} which is the map in the 2-disk on the left of Figure \ref{fig3}. It turns out that its nonwandering set consists of a sink and a hyperbolic homoclinic class containing the saddle. Its suspension is the flow described in the right-hand picture of the figure. It is a flow in the solid torus whose nonwandering set is also a periodic sink $O_1$ together with a hyperbolic homoclinic class. The next Figure \ref{fig4} describes a procedure of inserting singularities in the suspended Smale horseshoe. We select an horizontal interval $I$ and a point $x$ in the square forming the horseshoe. \begin{figure}[h] \begin{center} \input{vicosa21.pdf_t} \caption{Inserting singularities.}\label{fig4} \end{center} \end{figure} The selection is done in order to place $I$ in the stable manifold of a Lorenz-like equilibrium $\sigma_+$, and $x$ in the stable manifold of a Lorenz-like equilibrium for the reversed flow $\sigma_-$. This construction requires to add two additional singularities, a source $s_-$ to which the unstable branch of $\sigma_-$ not containig $x$ goes; and a saddle $s_+$ close to $\sigma_+$. See Figure \ref{fig0-5}. \begin{figure}[h] \begin{center} \input{vicosa1-21.pdf_t} \caption{Still inserting singularities.}\label{fig0-5} \end{center} \end{figure} An accurate description of the aforementioned procedure is done in \cite{bpv} and \cite{n}. Next we observe that the resulting flow's return map presents a cut along $I$ and a blowup circle derived from $x$. We now proceed to deform the flow in order to obtain a deformation of the return map by pushing up one branch of the circle, and pushing down the cusped region derived from the cutting as indicated in Figure \ref{fig5}. \begin{figure}[h] \begin{center} \input{vicosa31.pdf_t} \caption{Deforming.}\label{fig5} \end{center} \end{figure} We keep doing this deformation (see Figure \ref{fig6}) up to arrive to the final flow whose return map is described in Figure \ref{fig7}. \begin{figure}[h] \begin{center} \input{vicosa41.pdf_t} \caption{Still deforming.}\label{fig6} \end{center} \end{figure} \begin{figure}[h] \begin{center} \input{vicosa8-51.pdf_t} \caption{Return map.}\label{fig7} \end{center} \end{figure} This flow is defined in a solid torus, transversal to the boundary and pointing inward there. The final return map (denoted by $R$) is described with some detail in Figure \ref{fig8}. We are in position to describe the homoclinic classes $H_-$ and $H_+$ in Theorem \ref{thB}. They are precisely the maximal invariant set of $R$ in the upper and lower rectangles $Q_+$ and $Q_-$ forming the rectangle $Q$ in Figure \ref{fig8}. These maximal sets are located in the intersections $A\cap B\cap A'\cap B'$ (for $H_-$) and $C\cap D\cap D\cap E\cap C'\cap E'\cap D'$ (for $H_+$). A rough description of $H_-$ and $H_+$ is that $H_+$ is the singular horseshoe in \cite{lp} and $H_-$ its time reversal. \begin{figure}[h] \begin{center} \input{vicosa61.pdf_t} \caption{Localizing $H-$ and $H_+$ in $Q$.}\label{fig8} \end{center} \end{figure} The proof that $H_-$ and $H_+$ are nontrivial homoclinic classes is done as in \cite{b}, \cite{bmp}. The analysis in \cite{blmp} or \cite{lp} shows that $H_+$ is a sectional-hyperbolic set for the (final) flow and that $H_+$ is a sectional-hyperbolic set for the reversed flow. We assume that the horizontal conefield $x\in Q\mapsto C^s_1(x)$ and the vertical conefield $x\in Q\mapsto C^u_1(x)$, where $$ C^s_\alpha(x)=\left\{(a,b)\in\mathbb{R}^2:\frac{|b|}{|a|}\leq\alpha \right\} \mbox{ and } C^u_\alpha(x)=\left\{(a,b)\in\mathbb{R}^2: \frac{|a|}{|b|}\leq \alpha\right\}, \quad\forall \alpha>0, $$ are contracting and expanding (respectively) for the return map $R$ in the sense that there is $\rho>1$ with the following properties: \begin{enumerate} \item If $x\in R^{-1}(Q)\cap Q$ then $$ DR(x)C^u_1(x)\subset C^u_{\frac{1}{2}}(R(x)) \mbox{ and } \|DR(x)v^u\|\geq\rho\|v^u\|, \quad\forall v^u\in C^u_1(x). $$ \item If $x\in R(Q)\cap Q$ then $$ DR^{-1}(x)C^s_1(x)\subset C^s_{\frac{1}{2}}(R^{-1}(x)) \mbox{ and } \|DR^{-1}(x)v^s\|\geq \rho\|v^s\|, \quad\forall v^s\in C^s_1(x). $$ \end{enumerate} (See Figure \ref{fig8}.) Since such conefields do not allow the existence of nonhyperbolic periodic points, and are preserved by small perturbations, we obtain that the final flow is star in its solid torus domain. Next we observe that $H_-$ is not hyperbolic, since it contains the singularity $\sigma_-$ and, analogously, $H_+$ is not hyperbolic for it contains $\sigma_+$. Since every homoclinic class is transitive, we conclude from Theorem \ref{th1} that $H_-$ is not positively sectional-hyperbolic and $H_+$ is not negatively sectional-hyperbolic. To complete the proof we extend the final flow from its solid torus domain to the whole $S^3$. This is done by glueing it with another solid torus whose core is a periodic source $O_2$. This completes the proof. \qed
\section{Introduction} \IEEEPARstart{I}{n} human beings, sleep is a universal recurring dynamical and physiological activity, and the quality of sleep influences our daily lives in diverse ways. However, it was not until recently that sleep became a brach of medicine and found its role in several seemingly unrelated clinical problems. Physiologically, it is divided into two broad stages: rapid eye movement (REM), and non-rapid eye movement (NREM) \cite{Lee-Chiong:2008}. Normally, sleep proceeds in cycles in between REM and NREM. The NREM stage is further divided into shallow sleep (stage N1 and N2) and deep sleep (stage N3). In all procedures identifying sleep stages, we need a sleep scoring process with the help of polysomnography (PSG), which includes electroencephalography (EEG), electromyogram (EMG), and electrooculogram (EOG), etc. Among these physiological signals, EEG signals are the most concentrated ones since the clinically acceptable stage of the sleep is majorly determined by reading the recorded EEG based on the R\&K criteria, which were standardized in 1968 by Allan Rechtschaffen and Anthony Kales \cite{Rechtschaffen_Kales:1968} and further developed by the American Academy of Sleep Medicine on 2007 (AASM 2007) \cite{Iber_Ancoli-Isreal_Chesson_Quan:2007}. However, due to the subjective judgement and different training background, the agreement of manual sleep scoring among trained clinicians and professionals has been known to be limited \cite{Norman_Pal_Stewart_Walsleben_Rapoport:2000}, thereby motivating the development of an objective and automatic scoring system. Based on these clinical findings, various features of the EEG signals have been proposed to study the sleep dynamics, for example, time domain summary statistics, spectral analysis, coherence, time-frequency analysis, entropy, to name but a few \cite{Bajaj_Pachori:2013,Kannathal_Choo_Acharya_Sadasivan:2005,Blanco_Quiroga_Rosso_Kochen:1995,Geng_Zhou_Yuan_Cai_Zeng:2011}. Recently, a theoretically solid approach suitable to model the underlying dynamics of the brain activity and estimate the evolutionary dynamics from recorded EEG signal was proposed in \cite{TalmonPNAS,TalmonACHA}, and had been successfully applied to predict the pre-seizure state from the intra-cranial EEG signals \cite{Duncan_Talmon_Zaveri_Coifman:2013,TalmonTSP}. However, it is well known that sleep is a global and systematic behavior not localized solely in the brain. For example, the muscular atonia and low amplitude EMG are related to the significant changes in the breathing pattern during normal sleep: during NREM sleep, especially stage N3, breathing is remarkably regular, while during REM sleep, breathing is irregular with sudden changes. The above physiological facts hint that the respiratory pattern of the recorded breathing signal during sleep might well reflect the sleep stage. There have been some reported studies of the sleep stage from analyzing the respiratory signal \cite{Chung_Choi_Kim_Lim_Choi_Jeong_Park:2007,Guerrero-Mora_Elvia_Bianchi_Kortelainen:2012,Sloboda_Das:2011,Wu:2013,Chen_Cheng_Wu:2013}. In \cite{Chung_Choi_Kim_Lim_Choi_Jeong_Park:2007} (resp. \cite{Guerrero-Mora_Elvia_Bianchi_Kortelainen:2012}), an averaged respiratory rate over a fixed window is used to estimate the REM and NREM (resp. awake and sleep). In \cite{Sloboda_Das:2011}, a notch filter based instantaneous frequency estimator is applied to extract features to differentiate awake, REM and NREM. In \cite{Wu:2013,Chen_Cheng_Wu:2013}, the {\it adaptive harmonic model} and a modern time-frequency analysis technique have been applied to further quantify the notion of respiratory dynamic. In particular, the instantaneous respiratory rate has been related to awake, REM, shallow and deep sleep stages, with a rigorous mathematical foundation. The above-mentioned physiological patterns inside the EEG and the respiratory signals are actually outcomes of the intricate deformation of the underlying sleep dynamics, which we call {\it intrinsic dynamical features} of the sleep, that are not directly accessible to us. Although it is not an easy task to fully model or estimate the dynamical system underlying sleep, we might expect to benefit if we are able to quantify and integrate these hidden intrinsic dynamical features. In this paper, we propose to combine two modern adaptive signal processing techniques, {\it Empirical Intrinsic Geometry (EIG)} and {\it Synchrosqueezing transform (SST)}, to estimate these intrinsic dynamical features of sleep guiding the observed EEG and respiratory signals -- we define an index, referred to as {\it Sleep Index}, to quantify these features. Then, by applying the suitable classifier algorithm, we show that the extracted features are highly correlated to the sleep stage determined by reading the EEG by the AASM 2007 criteria. Indeed, the proposed classification based on the respiratory signal (resp. respiratory and EEG signals) has the overall accuracy $81.7\%$ (resp. $89.3\%$) in the relatively normal subject group, which is comparable to human expert classification. The article is organized in the following way. In Section \ref{Section:Method}, we summarize the theoretical background of EIG and SST and the associated models. Then the Sleep Index is introduced in Section \ref{Section:SleepIndex}. In Section \ref{Section:Result}, the proposed Sleep Index is applied to study the whole night sleep signals. We conclude with discussion in Section \ref{Section:Discussion}. \section{Two Algorithms -- Synchrosqueezing transform and Empirical Intrinsic Geometry}\label{Section:Method} The work presented in this paper is an application of the modern signal processing techniques to study the sleep dynamics. In particular, we will extract different features from the respiratory and EEG signals by the well studied EIG \cite{TalmonPNAS,Talmon_Cohen_Gannot_Coifman:2013} and SST \cite{Daubechies_Lu_Wu:2011,Chen_Cheng_Wu:2013}. As such, the theoretical material will be presented in a compact, informal manner emphasizing on the intuitions. We provide a formal and rigorous summary without proof of the details. Those interested in the proofs are encouraged to read the associated references. \subsection{Adaptive Harmonic Model and Synchrosqueezing Transform} The major characteristic pattern of the respiratory signal is that it is almost periodic. We call the movement of air from the environment into the lungs {\em inspiration} and the movement of air in the opposite direction {\em expiration}. An inspiration and an expiration constitute {\em a respiratory cycle}. {\em Breathing process} is a physiological process consists of a sequential respiratory cycles. In this paper, we focus on the breathing process and call the time-varying volume occupying the lung space {\em the physiological respiratory signal}. This general observation leads us to the following {\it phenomenological model} for the respiratory signal $R(t)$ (without noise): \begin{equation} R(t) = A(t) \, s(\phi(t)), \label{decomp1} \end{equation} where we shall call $s(\cdot)$ the wave shape function; it is a $1$-periodic real function that satisfies some mild technical conditions. See \cite{Wu:2013} for the details. The respiratory signals recorded from different devices, like the airflow measuring device or the chest wall movement, shall be understood as observations of the respiratory system. Different observations lead to different shape functions. We call the derivative $\phi'(t)$ of the function $\phi(t)$ the {\it instantaneous frequency} (IF) of the respiratory signal $R(t)$. We require IF to be positive, but it does not required to be constant as long as the variations are slight from one period to the next, i.e. $|\phi''(t)|\leq \epsilon\phi'(t)$ for all time $t$, where $\epsilon$ is some small, pre-assigned positive number. Likewise, We call $A(t)$ the {\it amplitude modulation} (AM) of $R(t)$, which should be positive, but is allowed to vary slightly as well, i.e. $|A'(t)|\leq \epsilon\phi'(t)$ for all time $t$. We refer the interested reader to \cite{Chen_Cheng_Wu:2013} for the technical details and a further discussion of the well-definedness of the definition of AM and IF. Note that our treatment of the respiratory signal is purely phenomenological; that is, the parameters and indices we will derive from the signal will be based solely on these signals themselves, and not on explicit, quantitative models of the underlying mechanisms. Physiologically, the quantities $\phi'(t)$, $A(t)$ and $s$ in the model (\ref{decomp1}) quantify the dynamics of the breathing process, which we refer to as {\it phenomenological dynamical features}. For example, one way to quantify the widely used notion {\it breathing rate variability (BRV)} \cite{Engoren:1998,Benchetrit:2000,Wysocki_Cracco_Teixeira_Mercat_Diehl_Lefort_Derenne_Similowski:2006} is considering the IF and AM \cite{Wu_Hseu_Bien_Kou_Daubechies:2013}. Indeed, if $\phi'(t_0)>\phi'(t_1)$ where $t_1\neq t_0$, we know that the subject breaths faster at time $t_0$ than at time $t_1$. We mention that this ``fast-slow'' momentary behavior in the respiratory signal has been shown to be clinically informative and can be helpful in the ventilator weaning prediction \cite{Wysocki_Cracco_Teixeira_Mercat_Diehl_Lefort_Derenne_Similowski:2006,Wu_Hseu_Bien_Kou_Daubechies:2013} and sleep stage estimation \cite{Wu:2013,Chen_Cheng_Wu:2013}. Due to the inevitable measurement error and other outliers appearing inside the system, we model the {\it recorded respiratory signal} as \begin{align}\label{observation_signal} Y(t) =R(t)+\sigma(t)\xi(t), \end{align} where $\xi(t)$ is a stationary generalized random process and $\sigma$ is a smooth function which varies slowly. Here $\xi(t)$ models the noise and other outliers and $\sigma$ models the possible non-stationary nature of the noise. A particular example for $\xi$ frequently encountered in practice is the Gaussian white noise. We refer the read having interest to \cite{Chen_Cheng_Wu:2013} for further information about this noise model and its mathematical details. To estimate the phenomenological dynamical features of $R(t)$, $\phi'(t)$ and $A(t)$, from $Y(t)$, we apply the {\it Synchrosqueezing transform} (SST), which is a special reallocation technique \cite{Daubechies_Lu_Wu:2011,Chen_Cheng_Wu:2013}. In a nutshell, we evaluate any linear time-frequency analysis on the observation $Y(t)$, for example the short time Fourier transform or the continuous wavelet transform, and we take the phase information hidden inside the chosen linear time-frequency analysis into account to obtain a sharpened time-frequency representation, which is denoted as $S_Y(t,\xi)$. In addition to capturing the phenomenological dynamical features of $R(t)$, the SST provides a sharper time-frequency representation compared with the other traditional time-frequency analyses. See Figure \ref{fig:1} for an example of the respiratory signal and its instantaneous frequency. We refer the reader to \cite{Daubechies_Lu_Wu:2011,Chen_Cheng_Wu:2013} for more details. \begin{figure}[h] \begin{centering} \includegraphics[width=.5\textwidth]{fig1_sleep_SST.png} \end{centering} \caption{The time-frequency representation of the respiratory signal determined by the Synchrosqueezing transform (SST) superimposed by the respiratory signal (the blue curve). The dominant black curve shown in the time-frequency representation indicated by the red arrows is the instantaneous frequency (IF) of the respiratory signal. It is clear that the subject breathes faster during the time indicated by the blue arrow than that indicated by the green arrow. This observation is captured by the IF indicated by the blue and green arrows.} \label{fig:1} \end{figure} \subsection{Empirical Intrinsic Geometry and its underlying Mathematical Model} \label{Subsec:EIG} In many real-world applications, the seeming complicated time series collected from the system is controlled by a relatively simple underlying process. In some situations, when the underlying evolutionary process lies on a low-dimensional Riemannian manifold, it can be parameterized through a manifold learning framework, which was first introduced and studied in \cite{Coifman_Singer:2008}. The main idea in \cite{Coifman_Singer:2008}\footnote{In the original paper \cite{Coifman_Singer:2008}, the method was referred to as Nonlinear Independent Component Analysis. However, for the sake of avoiding possible confusion with Independent Component Analysis, the name Empirical Intrinsic Geometry (EIG) was adpated in \cite{TalmonPNAS,TalmonACHA}.} and its extensions to time-series \cite{TalmonPNAS,TalmonACHA} is to bridge between data mining, and in particular manifold learning, and dynamical systems. The authors' observation that the accessible measurements at hand do not necessarily convey the true essence of the system led to the development of a more generalized model, which separates between measurements and intrinsic hidden variables. One particular example for such a dynamical system is the respiratory signal recorded during sleep -- we consider the model that the evolutionary process governing the respiratory signals is restricted to a low-dimensional Riemannian manifold, which is fundamentally different from the phenomenological model \eqref{observation_signal}. This dependency is encoded using the {\it state-space} formalism and the model of the recorded respiratory signal \eqref{observation_signal} is extended as follows: \begin{equation}\label{state_space} \left\{ \begin{array}{ll} Y(t) =R_{\boldsymbol{\theta}}(t)+\sigma(t)\xi(t),&\,\, \mbox{[measurement equation]}\\ \textup{d}\theta_i(t)=a_i(\theta_i(t))\textup{d} t+\textup{d} w_i(t),&\,\, \mbox{[state equation]} \end{array} \right. \end{equation} where $\boldsymbol{\theta}(t):=(\theta_1(t),\ldots,\theta_d(t))$ forms the {\it inaccessible intrinsic state} at time $t$ that governs the respiratory signal $R_\theta(t)$ and evolves in time with unknown drifts $a_i$ and independent standard Brownian motions $w_i$, $i=1,\ldots,d$. The idea that lies behind the model \eqref{state_space} is twofold. First, the measured signal $Y(t)$ has typical (unknown) dynamics, modeled here by the state equation, which need to be taken into account and encoded in the desired features. Second, the accessible signal is viewed as a measurement of the neural system controlling the sleep cycle. While it can be effected by numerous factors relating to the measurement modality (e.g., measurements of airflow or chest movements), the used equipment (e.g., the type of sensors and their exact positions), and noise, the true intrinsic variable we have interest in is the intrinsic states controlling the respiratory signal (represented here by $\boldsymbol{\theta} (t)$). Indeed, the notation $R_{\boldsymbol{\theta}}(t)$ implies that the respiratory signal depends upon the ``real" physiological variable $\boldsymbol{\theta}$ in an unknown way, possibly through its amplitude $A _\theta (t)$, wave shape $s _\theta (\cdot)$, or instantaneous frequency $\phi' _\theta (t)$. In the above model \eqref{state_space}, however, due to noise and other nuisance factors, the measurement $Y(t)$ might be too redundant to faithfully describe the dependency of the the respiratory signal on the underlying state and its temporal evolutionary. Thus, to improve the underlying state observability, we introduce a {\it high dimensional (possibly nonlinear) observer} $\Phi$ to the measured signal \cite{TalmonTSP}, i.e., \begin{equation}\label{equation:observation} \boldsymbol{Z}(t) = \Phi(Y(t)) \in \mathbb{R}^m, \quad \mbox{[observation equation]} \end{equation} where $\Phi$ is a map from the suitable scalar valued functional space to the $\mathbb{R}^m$-valued functional space and $m\geq1$ is an integer specified by the observer. With the sampled observation set $\mathcal{Z}:=\{\boldsymbol{Z}(t_i)\}_{i=1}^N$, a natural question is how to estimate $\boldsymbol{\theta}(t)$, namely, the system intrinsic state and dynamics of interest. Such an analysis may complement the phenomenological dynamical features provided by the SST. While the SST mainly carries instantaneous information, recovering the intrinsic state of the dynamical system $\boldsymbol{\theta} (t)$ provides a characterization of coarser, slower dynamical changes of the shape and structure of the signal, especially when the observer $\Phi$ is implemented as a transform that relies on short time frames analysis. It was shown in \cite{Coifman_Singer:2008} that if the observations $\boldsymbol{Z} (t)$ can be written as a regular deterministic function $f:\mathbb{R}^d \rightarrow \mathbb{R}^m$ of the samples of the underlying state, i.e., $\boldsymbol{Z}(t) = f(\boldsymbol{\theta} (t))$, then, by It\^o's formula, we have \begin{equation} \textup{d} Z_j(t)=\sum_{i=1}^d\left(\frac{1}{2}f^j_{ii}+a_i f^j_i\right)\textup{d} t + \sum_{i=1}^d f^j_i\textup{d} w_i(t) \end{equation} where $f^j_i=\partial f_j / \partial \theta _i$ and $f^j_{ii} = \partial ^2 f_j / \partial \theta _i ^2$. By a direct calculation, the covariance matrix $\boldsymbol{C}(t) \in \mathbb{R}^{m\times m}$ of the observation at time $t$ define by \begin{equation} C_{j,k}(t):=\text{Cov}(\textup{d} y_j(t),\textup{d} y_k(t)), \end{equation} satisfies $\boldsymbol{C}(t) =\boldsymbol{J} (t) \boldsymbol{J}^T(t)$, where $\boldsymbol{J}=\nabla f \in \mathbb{R}^{m\times d}$ is the Jacobian of $f$. This key result, along with the assumption that $\boldsymbol{\theta} (t)$ is locally stationary evolving much more slowly than the observation scale so that the it stays closely on a low-dimensional manifold $\mathcal{M}$ embedded in $\mathbb{R}^d$, which is referred to as the {\it intrinsic state manifold}, as well as the assumption that $f$ is stably invertible on its range, allow the authors in \cite{Coifman_Singer:2008} to estimate the inaccessible state through the solution of an eigenvector problem, which will be described later in this section. The main step leading to the solution theory is the following estimation. Suppose $\boldsymbol{\theta}(t),\boldsymbol{\theta}(\tau) \in \mathcal{M}$, $\boldsymbol{Z}(t)=f(\boldsymbol{\theta}(t))$ and $\boldsymbol{Z}(\tau)=f(\boldsymbol{\theta}(\tau))$. By the Taylor expansion of $f$, up to the error term $O(\|\boldsymbol{Z}(t)-\boldsymbol{Z}(\tau)\|^4)$ \cite{Coifman_Singer:2008}, we have: \begin{equation}\label{estimation:Mdist} \begin{array}{l} \|\boldsymbol{\theta} (t) - \boldsymbol{\theta} (\tau) \|_{\mathbb{R}^d}^2 =\frac{1}{2}(\boldsymbol{Z} (t) - \boldsymbol{Z}(\tau))^T \\ \quad \quad \quad \quad \quad \quad \times [\boldsymbol{C}^{-1}(t)+\boldsymbol{C}^{-1}(\tau)](\boldsymbol{Z} (t) - \boldsymbol{Z}(\tau)). \end{array} \end{equation} Note that in our example, the function $f$ leading to the observation depends on the observer $\Phi$. As a result, with the estimated covariance matrix from the accessible collected data $\mathcal{Z}$, we can build a {\it graph Laplacian} associated with the intrinsic state manifold from the finite observations $\mathcal{Z}$ using the estimated Euclidean distance between the corresponding underlying samples $\boldsymbol{\theta} (t_i)$ (\ref{estimation:Mdist}). This graph Laplacian gives rise to re-parametrization of the intrinsic manifold through diffusion maps (DM) \cite{Coifman_Lafon:2006}. This re-parametrization procedure aiming to extract the intrinsic dynamics of the observation is referred to as Empirical Intrinsic Geometry (EIG). The remaining key question is the choice or design of a ``proper" observer $\Phi$ in (\ref{equation:observation}) to the system. In particular, in order to accommodate the inevitable noise in real-world signals, estimates of the conditional probability density $p(\boldsymbol{Z} | \boldsymbol{\theta})$ (e.g., histograms) were proposed as observers in \cite{TalmonPNAS}. The analysis relies on the following facts: (a) any measurement noise is translated to a linear transformation in the conditional densities domain, and (b) the distance (\ref{estimation:Mdist}), which is the {\it Mahalanobis distance}, is invariant under linear transformations. Indeed, these two facts allow for the estimation of the distance between two nearby samples on the intrinsic manifold in adverse noisy conditions. However, estimating the conditional probability densities requires a large amount of data and often is not feasible. Unfortunately, standard representations based on the Fourier transform are also inadequate for respiratory signals. By linear approximation of the function $\phi(t)$ around a nearby sample at $t_0$, the respiratory signal in \eqref{decomp1} can be approximated by \begin{equation} R(t) \approx A(t_0) \, s(\phi(t_0) - t_0 + \phi'(t_0)t), \end{equation} As a result, the modulus of the Fourier transform of $R(t)$ around $t_0$ is approximated by \begin{equation}\label{time_deformation} |\hat{R}(t_0, \omega) | \approx |A(t_0) \, \hat{s}(\omega / \phi'(t_0))|, \end{equation} where $\omega$ is the frequency, $\hat{R}(t_0, \omega)$ is the Fourier transforms of $R(t)$ around $t_0$ and $\hat{s}(\omega)$ is the Fourier transform of $s(t)$, respectively, assuming $A(t)$ changes slowly with time. The approximation in \eqref{time_deformation} implies that even an almost linear function $\phi(t)$ (i.e., when the IF is $\phi'(t) \approx 1$) is translated to large deformations in the Fourier domain in high frequencies \cite{Mallat2012}. Consequently, if we take the short time Fourier transform as the observer, the observations exhibit instabilities, thereby leading to irregular $f$ and a poor estimation of the Euclidean distance on the underlying intrinsic state manifold \eqref{estimation:Mdist}. To overcome the instability of the Fourier representation, following \cite{TalmonTSP}, we use the scattering transform as an observer. The scattering transform has a low variance because it is based on first order moments of contractive operators, it linearizes deformations, and it can represent effectively intermittent behavior \cite{Mallat2012,Bruna2013}. The scattering transform is computed based on a cascade of wavelet transforms and nonlinear modulus operators \cite{Mallat2012}. Here, we briefly review the construction procedure of its first and second order levels, since they were empirically shown to provide a sufficient representation of the signals considered in this paper. Let $\psi (t)$ be a complex wavelet, whose real and imaginary parts are orthogonal and have the same $L_2$ norm. Let $\psi _j (t)$ denote the dilated wavelet, defined as $\psi _j (t) := 2^{-j} \psi (2^{-j}t), \ \forall j \in \mathbb{Z}$. Let $\Phi _s (R_{\boldsymbol{\theta}}(t))$ denote the observations computed by applying the first and second level scattering transform to the signal samples $R_{\boldsymbol{\theta}}(t)$, which are given by \begin{align*} &\Phi _s (R_{\boldsymbol{\theta}}(t)) = (|| R_{\boldsymbol{\theta}}(t) * \psi _{j_1} (t) | * \psi _{j_2}(t)| * w (t) \\ &\qquad\qquad\qquad \forall (j_1, j_2) \in \mathbb{Z}^n, n \in \{1,2\})_{j_1,j_2} \end{align*} where $w(t)$ is a smoothing window, i.e., a scaling function associated with the mother wavelet. The scattering transform has been shown to be an observer that is especially suitable for deformations and intermittencies \cite{TalmonTSP}. In particular, it was shown that it is regular with respect to time deformations. Therefore, the application of the scattering transform to the respiratory signal is particularly suitable. Building on the generality of the described analysis, in this study, we use it to represent the EEG signals as well. As the respiratory signal, the EEG signal measures a physiological phenomenon (``the brain activity"), but, it is subject to noise, interferences, and nuisance factors. Likewise, it can be represented using a state-space model, similar to \eqref{state_space}, given by \begin{equation* \left\{ \begin{array}{ll} X(t) = E_\zeta(t) + V(t)\,&\,\, \mbox{[measurement equation]}\\ \textup{d}\zeta_i(t)=\alpha_i(\zeta_i(t))\textup{d} t+\textup{d} u_i(t),&\,\, \mbox{[state equation]} \end{array} \right. \end{equation*} where $E_{\zeta}(t)\in\mathbb{R}^{m'}$ and $X(t)\in\mathbb{R}^{m'}$ are the clean and noisy EEG signals, $V(t)$ is a measurement noise, and $\boldsymbol{\zeta}(t):=(\zeta_1(t),\ldots,\zeta_{d'}(t))$ denotes the inaccessible intrinsic state representing the brain activity that governs the EEG signal $E_\zeta(t)$ and evolves in time with unknown drifts $\alpha_i$ and independent standard Brownian motions $u_i$, $i=1,\ldots,d'$. By applying EIG to the recorded EEG signals, we may reconstruct the intrinsic states $\boldsymbol{\zeta}$. We remark that this approach was applied to identify the pre-seizure state from intracranial EEG data \cite{Duncan_Talmon_Zaveri_Coifman:2013,TalmonTSP}. We refer the interested reader to \cite{TalmonPNAS,TalmonACHA,TalmonTSP} for more technical details and references. Before closing this section, we summarize the construction of the graph Laplacian parametrization. In a nutshell, the main ingredient is integrating local similarities at different scales, which leads to a global description of the data set. Unlike linear methods such as principal component analysis (PCA), a graph Laplacian parametrization embodies nonlinear relationships among the variables. In addition to the mathematical analysis results \cite{Coifman_Lafon:2006,Belkin_Niyogi:2007,Singer_Wu:2013}, it has been shown to be robust to noise perturbation \cite{ElKaroui:2010a,ElKaroui_Wu:2014} and it is computationally efficient. We outline the algorithm here and refer the readers to these literatures for further theoretical details. Take $N$ multivariate measurement samples $\mathcal{Z}=\{\boldsymbol{Z}(t_i)\}_{i=1}^N$ and build a complete graph with vertices $\mathcal{Z}$. We first build an affinity matrix (or adjacency matrix) $\mathsf{W}$ of size $N \times N$. The affinity between a pair of samples is defined by a metric $d$ in the following way: \begin{equation} \label{W} W_{ij}=e^{-\frac{d^2(\boldsymbol{Z}(t_i),\boldsymbol{Z}(t_j))}{\epsilon}}, \quad \mbox{for } i,j=1,\ldots,N,\, i\neq j. \end{equation} Note that according to the noise analysis in \cite{ElKaroui_Wu:2014}, when the signal to noise ratio is small, it is beneficial to set the diagonal terms of the affinity matrix to $0$. In the present work, following the analysis in \cite{Coifman_Singer:2008}, the metric we choose is the Mahalanobis distance \eqref{estimation:Mdist}. It is clear that the matrix $\mathsf{W}$ is symmetric. Note that theoretically (and practically) we can choose a more general kernel function, but we focus on the Gaussian kernel to simplify the exposition. Then we define the diagonal degree/density matrix $\mathsf{D}$ of size $n\times n$, consisting of the sum of rows of $\mathsf{W}$: $$ D_{ii}=\sum^N_{j=1}W_{ij},\quad \mbox{for } i=1,\ldots,N. $$ Based on $\mathsf{W}$ and $\mathsf{D}$, the {\it graph Laplacian} is defined by $$ \mathsf{L}:=\mathsf{I}-\mathsf{D}^{-1} \mathsf{W}. $$ Note that under the manifold assumption, $\mathsf{D}^{-1}$ exists. Also note that $\mathsf{D}^{-1}\mathsf{W}$ can be viewed as a transition matrix of a Markov chain on the samples. Since $\mathsf{L}$ is similar to the symmetric matrix $\mathsf{I}- \mathsf{D}^{-1/2} \mathsf{W} \mathsf{D}^{-1/2}$, it has a complete set of right eigenvectors $\varphi_1,\varphi_2,\ldots,\varphi_{N}$ with corresponding eigenvalues $0=\lambda_1<\lambda_2\leq\dots\leq\lambda_{N}\leq1$, where $\varphi_1=(1,1\dots,1)^T$ \cite{Coifman_Lafon:2006}. By the above construction, the eigenvectors $\varphi_1,\ldots,\varphi_{N}$ are vectors in $\mathbb{R}^N$. Through the eigenvectors, the measurement samples are mapped into $\mathbb{R}^{\hat{d}}$ via \begin{equation} \label{d-map} \boldsymbol{Z}(t_i) \mapsto (\varphi_2(t_i),\ldots,\varphi_{\hat{d}}(t_i)),\quad \mbox{for } i=1,\ldots,N. \end{equation} where $\hat{d}$ is an estimate of the dimension of the intrinsic state of the system and is usually $\hat{d}\ll N$. Estimating the intrinsic dimension of the system $d$ extends the scope of the paper and is empirically set according to the spectral gap in the decay of the eigenvalues, as will be described in Section \ref{Section:SleepIndex}. In \eqref{d-map}, we obtain a $\hat{d}$-dimensional parameterization of the measurements. In particular, we view the $j$th coordinate of the parameterization of $\boldsymbol{Z} (t_i)$, i.e., $\varphi_{j+1} (t_i)$, as the $j$th coordinate of the recovered hidden intrinsic state $\theta _j (t_i)$, which we view as the features associated with the sleep stage in this analysis. An illustration of the DM reparametrization process with the first $3$ non-trivial eigenvectors is shown in Figure \ref{fig:2}. \begin{figure*}[t] \begin{centering} \includegraphics[width=.5\textwidth]{Figure2_rt.png} \includegraphics[width=.5\textwidth]{fig2_O1A2_NLICA.png} \end{centering} \caption{The intrinsic dynamical features of the cortical activity extracted from the O1A2 EEG signal by the scattering Empirical Intrinsic Geometry (EIG). On the left, the scattering EIG is illustrated -- the graph Laplacian is built up from the Mahalanobis distance from the EEG signal via the scattering operator. On the right, the top three nontrivial eigenvalues of the graph Laplacian are used to show the underlying evolutionary dynamics. The blue circles (resp. cyanid circles, yellow circles and red circles) represent the awake (resp. REM, N1 and N2 and N3) sleep stage. It is clear that the extracted dynamical features well parametrize the sleep stages in the sense that different sleep stages are located in different places.} \label{fig:2} \end{figure*} \section{Material and Method}\label{Section:SleepIndex} \subsection{Data Collection} Standard polysomnography was performed with at least 6 hours of sleep to confirm the presence or absence of OSA from the clinical subjects suspicious of sleep apnea in the sleep center at Chang Gung Memorial Hospital (CGMH), Linkou, Taoyuan, Taiwan. The institutional review board of the CGMH approved the study protocol (No. 101-4968A3) and the enrolled subjects provided written informed consent. Four channel EEG signals (C3A2, C4A1, O1A2 and O2A1), two channel EOG signals and chin EMG were recorded at the sampling rate $200$ Hz for sleep staging. Chest and abdominal motions are recorded by the piezo-electric bands and airflow was measured using thermistors and nasal pressure, both at the sampling rate $100$ Hz. All signals were acquired on the Alice 5 data acquisition system (Philips Respironics, Murrysville, PA). Apneas and hypopneas were defined using AASM 2007 guidelines \cite{Iber_Ancoli-Isreal_Chesson_Quan:2007}, and the apnea-hyponea index (AHI) provided is the value determined during sleep. Take the recorded EEG signals, denoted as $E_k$, $k=1,\ldots,4$, and the respiratory signal, denoted as $R$. Suppose the recording time period is $\mathcal{T}=[0,T]$. We divide $\mathcal{T}$ into contiguous subintervals $\mathcal{T}_i$ of $\tau$ seconds long, $i=1,\ldots,N$; that is, $\mathcal{T}=\cup_{i=1}^N \mathcal{T}_i$ and $\mathcal{T}_i\cap \mathcal{T}_j=\emptyset$ for all $i\neq j$. We call $\mathcal{T}_i$ the {\em $i$-th epoch}. We will extract $p>0$ features out of the recorded respiratory and EEG signals for each epoch. \subsection{Features from the respiratory signal} Given a recorded respiratory signal $R(t)$, we extract its {\it phenomenological dynamical features}, including the instantaneous frequency $\phi'(t)$ and the amplitude modulation $A(t)$ by applying the SST. Denote the estimated instantaneous frequency by $\tilde{\phi}'(t)$ and the amplitude modulation by $\tilde{A}(t)$. The mean of $\tilde{A}(t)$ restricted to the $i$-th epoch, denoted as the $A_i$, and the mean of $\tilde{\phi}'(t)$ restricted to the $i$-th epoch, denoted as $\phi'_i$, form the first two features for the respiratory signal for the $i$-th subinterval. The third feature, denoted as $v_i$, is obtained by evaluating the standard deviation of $\tilde{\phi}'(t)$ on the interval of length $30$ seconds centered on the middle of the $i$-th epoch. We apply the analysis described in Section \ref{Subsec:EIG} to $R(t)$ in order to complement the phenomenological dynamical features and to obtain a characterization of the structural, slower underlying variables of the data. Here as well we obtain the graph Laplacian $\mathsf{L}^{(R)}\in \mathbb{R}^{N\times N}$. Then, the eigenvectors and eigenvalues of $\mathsf{L}^{(R)}$ are given by $\mathsf{L}^{(R)} \varphi_{j}^{(R)}=\lambda_{j}^{(R)}\varphi_{j}^{(R)}$. The first $\hat{d}^{(R)} \geq 1$ nontrivial eigenvectors are chosen based on the following ``spectral gap" thresholding criteria \begin{align}\label{equation:thesholding} \lambda_{\hat{d}^{(R)}+1 }^{(R)} <\delta \quad\mbox{ and }\quad \lambda_{\hat{d}^{(R)} +2}^{(R)} \geq\delta, \end{align} where $0<\delta<1$ is the threshold chosen by the user. Thus, using \eqref{d-map}, we obtain $\hat{d}^{(R)}$ {\it intrinsic dynamical features} of the respiratory system. \subsection{Features from the EEG signal} Given the EEG signal $E_k(t)$ recorded from the $k$-th channel, we run the analysis described in Section \ref{Subsec:EIG} and obtain the graph Laplacian $\mathsf{L}^{(E,k)} \in \mathbb{R}^{N\times N}$. Then, the eigenvectors and eigenvalues of $\mathsf{L}^{(E,k)} $ are given by $\mathsf{L}^{(E,k)} \varphi_{j}^{(E,k)}=\lambda_{j}^{(E,k)} \varphi_{j}^{(E,k)}$ with $0=\lambda_{1}^{(E,k)}\leq\lambda_{2}^{(E,k)}\leq\ldots$. The first $\hat{d}^{(E,k)}\geq 1$ nontrivial eigenvectors are chosen based on the thresholding criteria (\ref{equation:thesholding}) with the same $\delta$. Using the eigenvectors, each subinterval of the EEG signal $E_k(t)$ is mapped into a sub-vector of $\hat{d}^{(E,k)}$ dimensions according to \eqref{d-map}. By collecting the low dimensional vectors of all the channel, we obtain a vector consisting of $\sum_{k=1}^4 \hat{d}^{(E,k)}$ {\it intrinsic dynamical features of the cortical activity} for each subinterval. \subsection{Sleep Index} We consider the following two feature vectors. The first one is extracted only from the respiratory signal and is referred as the {\it Respiratory Index}: \begin{align*} \mathsf{r}_i:=\big(A_i,\phi'_i,v_i,\varphi^{(R)}_{2}(i),\ldots,\varphi^{(R)}_{\hat{d}^{(R)}+1}(i)\big). \end{align*} The second one is extracted only from the EEG signals and is referred as the {\it EEG Index}: \[ \mathsf{e}_i:=\big(\varphi_{2}^{(E,1)}(i),\ldots,\varphi^{(E,1)}_{\hat{d}^{(E,1)}+1 }(i),\ldots,\varphi^{(E,4)}_{2}(i),\varphi^{(E,4)}_{\hat{d}^{(E,4)}+1 }(i)\big). \] An analysis result of the O1A2 EEG signal, denoted as $\big(\varphi_{2}^{(E,1)}(i),\ldots,\varphi^{(E,1)}_{4}(i)\big)$, is shown in Figure \ref{fig:2}. Clearly different sleep stages represented in different colors have different features and are well clustered. In addition, these different sleep stages are organized in a continuous but nonlinear way -- from the right hand side of the figure to the left hand side we have awake, REM, N1 and N2 and deep sleep stages. Next, the $3$ phenomenological respiratory features, the $\hat{d}^{(R)}$ intrinsic respiratory features and the intrinsic dynamical features of the cortical activity at the $i$-th epoch are combined together to comprise the {\it Sleep Index} with $p=\sum_{k=1}^4\hat{d}^{(E,k)}+3+\hat{d}^{(R)}$: \[ \mathsf{s}_i:=\big(\mathsf{r}_i,\mathsf{e}_i\big). \] \subsection{Sleep Stage Classifier} Support vector machine (SVM) is a commonly used technique for the purpose of classification in statistical learning theory \cite{Scholkopf_Smola:2002}. In a nutshell, SVM determines a hyperplane in the space separating the data set into two disjoint subsets, such that each subset is lying in a different side of the hyperplane. With the help of the reproducing kernel Hilbert space theory, SVM is generalized to the {\it kernel SVM}, which allows for classification with nonlinear relationship; that is, a nonlinear surface separating the data set into two disjoints subsets may be used. We refer the interested reader to \cite{Scholkopf_Smola:2002} for technical details. For the sake of identifying the (possible) nonlinear relationship between different sleep stages, in this work we choose the radial based function (RBF), $K(\boldsymbol{x},\boldsymbol{x}')=\exp(-\frac{\|\boldsymbol{x}-\boldsymbol{x}'\|^2_2}{2\sigma^2})$, where $\sigma>0$, as the kernel function. Note that our dataset is multi-class -- the response has more than $2$ categories -- therefore, we need to further generalize the kernel SVM to the multi-class SVM to complete our mission. To this end, we apply the one versus all (OVA) classification scheme \cite{Rifkin_Klautau:2004}. Despite its simplicity, the OVA classification scheme is highly effective and useful, as was extensively shown and discussed in \cite{Rifkin_Klautau:2004}. Group data will be reported as mean $\pm$ standard deviation unless otherwise specified. \section{Result}\label{Section:Result} Ten subjects without sleep apnea (AHI less than $5$) were chosen for this study. The demographic characteristics of the individuals whose data was used are as follows: $6$ males and $4$ females, age: $45.9 \pm 12.3$ years, range $28-61$ years; BMI: $23.6\pm 1.9 \text{kg/m}^2$, range $21.5-28 \text{kg/m}^2$ ; AHI: $1.9 \pm 1.1$, range $0.4-3.4$. The total recorded time are of length $384\pm 27.8$ minutes with range $363-443$ minutes and we have a sleep period time of $367\pm 27.5$ minutes with range $338-428$ minutes for the sleep stage estimation. We divide the whole night sleep into contiguous epochs of length $2.56$ seconds. We take $\delta=0.01$ and the dimension of the Sleep Index $p$ is $11.2\pm1.69$. We consider the sleep stages in this study: \begin{align*} \mathcal{R} &=\{\text{Awake},\,\text{REM},\,\text{N1},\text{N2},\,\text{N3}\}=:\{\texttt{1},\texttt{2},\ldots,\texttt{5}\}. \end{align*} Here to simplify the notation, we reindex the set of sleep stages and use the teletype-font to avoid confusion; that is, $\texttt{1}$ is the awake stage, etc. Then we generate the different indices, $\{\mathsf{s}_i\}_{i=1}^N$, $\{\mathsf{r}_i\}_{i=1}^N$ and $\{\mathsf{e}_i\}_{i=1}^N$, from the recorded EEG and respiratory signals. The sleep stages in $\mathcal{R}$ are determined by the sleep expert as the ground truth. The OVA kernel SVM with the RBF kernel with $\sigma=1$ is applied to classify the different sleep stages. Suppose there are $n_\ell$ subintervals with sleep stage $\ell$, where $\ell=\texttt{1},\ldots,\texttt{5}$, in the validation dataset. Denote $n_{i,j}$ to be the number of subintervals with the sleep stage $\texttt{i}$ as the gold standard, but classified as the sleep stage $\texttt{j}$. We call the $5\times 5$ matrix $N$ with the $(i,j)$-th entry $n_{i,j}$ the {\it confusion matrix}. We also define the {\it confusion percentage matrix} $P$ as a $5\times 5$ matrix with its $(i,j)$ entry $\frac{n_{i,j}}{\sum_{j=1}^5n_{i,j}}$. We will call $P_{i,i}$ the {\it sensitivity (SE)} of the sleep stage $\texttt{i}$ prediction, which is denoted as SE($\texttt{i}$). We will also report the overall accuracy (AC) denoted as $\text{AC} :=\frac{\sum_{\texttt{i}=\texttt{1}}^{\texttt{5}}n_{\texttt{i},\texttt{i}}}{\sum_{\texttt{i,j}=\texttt{1}}^{\texttt{5}}n_{\texttt{i},\texttt{j}}}$ and the specificity (SP) of the sleep stage $\texttt{i}$ denoted as $\text{SP}(\texttt{i}):=\frac{ \sum_{\texttt{j}\neq \texttt{i}}n_{\texttt{j},\texttt{j}}}{\sum_{\texttt{k}}\sum_{\texttt{j}\neq \texttt{i}}n_{\texttt{j},\texttt{k}}}$. Note that these definitions are direct generalizations of the AC, SE and SP of the binary categorical response data. To prevent over-fitting and confirm the classification result, we run the repeated random sub-sampling validation $25$ times and evaluate the average. To be more precise, we randomly partition the data into the training dataset and the validation dataset -- the training dataset comprises $80\%$ of the features and the rest are used to form the validation dataset. The trained classifier based on the training dataset is applied to predict the sleep stages of the validation dataset. With the above preparation, first, we show that the proposed features capturing the sleep information hidden inside the respiratory signal are not only theoretically rigorously supported, but also useful in practice. The overall AC is $81.7\%$. The error bar of SE and SP of correlating the Respiratory Index and the sleep stages $\mathcal{R}$ over $25$ repeated random sub-sampling validation for the $10$ subjects is shown in the light gray curve in Figure \ref{fig:errorbar}. The average SE's (resp. SP's) over $10$ subjects for the awake, REM, N1, N2 and N3 stages are $82\%$, $89\%$, $72\%$, $82\%$ and $62\%$ (resp. $81\%$, $81\%$, $83\%$, $82\%$ and $82\%$). Second, we show that the EEG Index also correlates with the sleep stages. The overall AC is $71.6\%$. The error bar of the SE and SP over $25$ repeated random sub-sampling validation for the $10$ subjects is shown in the dark gray curve in Figure \ref{fig:errorbar}. The average SE's (resp. SP's) over $10$ subjects for the awake, REM, N1, N2 and N3 stages are $70\%$, $67\%$, $50\%$, $74\%$ and $54\%$ (resp. $70\%$, $71\%$, $73\%$, $65\%$ and $71\%$). Next, we combine all the features extracted from the respiratory signal and the EEG signals and show the result is better than simply using the Respiratory Indices or EEG Indices. The overall AC is $89.3\%$. The error bar of the SE and SP over $25$ repeated random sub-sampling validation for the $10$ subjects is shown in the black curve in Figure \ref{fig:errorbar}. The average SE's (resp. SP's) over $10$ subjects for the awake, REM, N1, N2 and N3 stages are $85\%$, $94\%$, $79\%$, $90\%$ and $68\%$ (resp. $89\%$ $89\%$, $90\%$, $87\%$ and $90\%$). We then apply the Mann-Whitney U test to test if the Sleep Index better predicts sleep stage than the Respiratory Index under our setup. The p value less than $0.01$ is considered significant. For the $25$ realizations of sub-sampling validation from $10$ subjects, we obtained $250$ SE's and $250$ SP's for different indices respectively. The Mann-Whitney U test is applied to see if the SE's and SP's are significantly different. The performance of the Sleep Index compared with the Respiratory Index on the awake, REM, N1, N2 and N3 stages in the sense of SE (resp. SP) are all significant with p-values $<0.001$ ($<0.001$). \begin{figure*}[t] \begin{centering} \includegraphics[width=.7\textwidth]{SE10.png} \end{centering} \caption{The error bar of the performance of each features for predicting the sleep stage. The upper (resp. lower) subfigure is the sensitivity (resp. specificity) of predicting different sleep stages by different indices over $25$ repeated random sub-sampling validation. The black (resp. light gray and dark gray) curve is for the Sleep Index (resp. Respiratory Index and EEG Index). The x-axis is the subject index ranging from $1$ to $10$.} \label{fig:errorbar} \end{figure*} Lastly, to better present the classification result, the averaged confusion percentage matrices over all subjects and sub-sampling realizations based on the Respiratory Index, EEG Index and the Sleep Index are shown in Figure \ref{fig:confusion}. Note that the diagonal entries are the SE's of sleep stage prediction. \begin{figure*}[t] \begin{centering} \includegraphics[width=.7\textwidth]{Confusion.png} \end{centering} \caption{The averaged confusion percentage matrices over all subjects and sub-sampling realizations based on the Respiratory Index (resp. EEG Index and Sleep Index) is shown in the left (resp. middle and right) subfigure. The percentage is represented by the color. The darker the entry is, the higher the value is. The precise value is shown in the color bar on the top of each matrix. Here, \texttt{1} (resp. \texttt{2}, \texttt{3}, \texttt{4} and \texttt{5}) in the x- and y-axis tick label stands for awake (resp. REM, N1, N2 and N3). It is clear to see the inclination of mis-classifying N3 into N2.} \label{fig:confusion} \end{figure*} \section{Discussion}\label{Section:Discussion} The results in Section \ref{Section:Result} show that an accurate estimation of all sleep stages by solely analyzing the respiratory signal is possible by combining EIG and SST. Indeed, in addition to the overall AC $81.7\%$, the average SE is greater than $72\%$ except N3, and the average SP is greater than $81\%$. On the other hand, we mention that while the features of the respiratory signal extracted by EIG and SST are complementary, only EIG can be applied to the EEG signal, since the EEG signal can not be modeled by the adaptive harmonic model. The overall AC based on EIG applied to the EEG signals is $71.8\%$, the SE is greater than $67\%$, except N1 and N3, and the average SP is greater than $65\%$. Namely, the performance of the EEG Index is not better than the Respiratory Index. Nevertheless, we see an improvement in the classification performance based on the Sleep Index, which contains information from both the respiratory and EEG signals. The overall AC is increased to $89.3\%$, the average SE is now greater than $79\%$ except N3, and the average SP is greater than $87\%$. In addition, it has been shown that the SE and SP of the Sleep Index are significantly better than those of the Respiratory Index. Moreover, the confusion percentage matrices also indicate that except N3, the mis-classification does not land in any specific sleep stage. The above findings lead to the following two tentative conclusions: 1. in addition to the EEG signals, the respiratory signal contains ample information about the sleep stage; 2. combining the relevant but different information hidden inside the respiratory and the EEG signals leads to a better result. The main innovation in our sleep depth analysis is the combination of the clinical observation and modern adaptive signal processing techniques. From the clinical standpoint, we take the well known physiological fact that in addition to the brain activity, sleep is a global dynamical process involving different sub-system dynamics, in particular the significant changes in the respiratory pattern among different sleep stages. From the signal processing standpoint, we emphasize the importance of the nonlinearity controlling the sleep cycle and focus on finding suitable mathematical tools not only adaptive to the signal but also with sufficient rigorousness to quantify the clinical observation. Indeed, since the unaccessible intrinsic sleep dynamics is reflected in the nonlinear behavior of the respiration, and the two modern signal processing techniques, EIG and SST, have being theoretically studied to well quantify these nonlinearity, we obtain effective features by analyzing the recorded respiratory signal, which surrogate the intrinsic sleep dynamics. The meaning of accuracy deserves some discussion. It is well known that the sleep stage determination agreement between different sleep experts is limited to $85\%$ even when the subjects under examination are normal, and it is even worse on the abnormal subjects \cite{Norman_Pal_Stewart_Walsleben_Rapoport:2000}\footnote{It is reported in \cite{Norman_Pal_Stewart_Walsleben_Rapoport:2000} that the mean agreement in the normal subset is higher (mean 76\%, range 65-85\%) than in the subset of sleep disordered breathing (mean 71\%, range 65-78\%).}. Although our cases are not diagnosed as sleep apnea, they cannot be considered as in the normal population either, thereby attaining accuracy rates higher than $80\%$ in our subjects may not be meaningful. On the other hand, we found that the classification of N3 stage is consistently worse and its mis-classification tends to land in N2, as is shown in Figure \ref{fig:confusion}. Notice that the subjects in our study are on average $48$ years old, and the distribution of N3 sleep stages in the normal population of this age is $4-20\%$. However, the N3 sleep stages in our study cases is $3.1\%\pm 3.26\%$ with $25\%$ and $75\%$ quantiles $0.8\%$ and $5\%$ respectively, which is much fewer than those in the normal population. Since the number of N3 in the training set is relatively small, even by applying the weighted SVM to handle the unbalanced data, we do not expect to attain a compatible classification rate of N3. This unbalanced training set issue, combined with the stable breathing pattern during N2 and N3, might explain the inclination of mis-classifying N3 into N2. Furthermore, while the accuracy of our classification is compatible with/better than the state-or-art reported results, we are able better classify between different sleep stages. Indeed, in \cite{Guerrero-Mora_Elvia_Bianchi_Kortelainen:2012}, the overall accuracy of classifying awake and sleep is $83.6\%$ based on the respiratory signal; in \cite{Chung_Choi_Kim_Lim_Choi_Jeong_Park:2007} an averaged respiratory rate classifies REM and NREM with accuracy over $85\%$; in \cite{Sloboda_Das:2011}, a notch filter based IF estimator is applied to extract respiratory features, which classifies awake, REM and NREM with mean accuracy approximately $70\%$; in \cite{Chen_Cheng_Wu:2013}, the IF estimated by SST is shown to be able to distinguish awake, REM, shallow and N3 with statistical significance. With the above discussions, we conclude that our features and the selected classifier are accurate. The sleep depth estimation by the EEG Index is inferior with respect to the traditional EEG analysis. To understand this result, we briefly revisit how a sleep expert determines the sleep stage. Based on the protocol criteria, in addition to an EEG signal of duration that exceeds 30 seconds, the expert also takes into account past and future EEG signals to determine the sleep stage. However, in our study, the EEG Index is based on the signal in epochs of length $2.56$ seconds. The choice of $2.56$-second interval is for the sake of balancing between the dimension and number of data points in EIG. Although the local covariance structure of the EEG signal is taken into account in the EIG analysis, this relationship is different from the protocol criteria. As a result, we do not expect to obtain a compatible stratification power. However, we see that even if we only focus on these short-term EEG signals, we still can predict the sleep stage up to some accuracy and it does help to attain a better classification rate when combined with the Respiratory Index. This hints the possibility that some useful information is hidden inside a finer scale EEG signals. This interesting potential will be reported in the future study. The discussion would not be complete without mentioning the shortcomings of our study. First, we focus on a small database containing only $10$ relatively normal subjects in this study. To confirm the usefulness of the proposed features, we need to study a larger database with different types of subjects. Second, the chosen features, in particular the features selected by EIG, are subject-dependent. Indeed, different subjects might have different dynamical systems and the number of dominant components determined by EIG might vary. A theoretical and practical study of integrating the proposed features among different subjects is undergoing. In conclusion, by applying modern signal processing techniques to EEG and respiratory signals, we find a set of suitable features, which allow us to predict the sleep stages accurately. In addition to gaining insight into the dynamics controlling the sleep dynamics, the automatic annotation system based on the analysis might lead to an objective classification as well as reduce the required human expert analysis involved in sleep evaluation. \section*{Acknowledgements} Hau-tieng Wu and Ronen Talmon thank the helpful discussions with Professor Ronald Coifman. Ronen Talmon acknowledges the support by the European Union's - Seventh Framework Programme (FP7) under Marie Curie Grant No. 630657. Yu-Lun Lo acknowledges the support by Taiwan National Science Council grant 101-2220-E-182A-001 and 102-2220-E-182A-001. \bibliographystyle{IEEEtran}
\section{Introduction} The dynamical instability of the astrophysical objects is the subject of interest in classical physical as well as in general theory of relativity (GR). The motivation of this problem become important when static stellar models are stable against the fluctuations produced by the self gravitational attraction of the massive stars. It is most relevant to structure formation during the different phases of the gravitationally collapsing objects. In the relativistic astrophysics the dynamical stability of the stars was studied by the Chandrasekhar \cite{1} in 1964 , since then a renowned interest has grown in this research area. Herrera et al.\cite{2,3} have extended the pioneers work for non-adiabatic, anisotropic and viscous fluids. All these investigations imply that adiabatic index ${\Gamma}_1$ define the range of instability, for example for the Newtonian perfect fluid such range is ${\Gamma}_1<4/3$. Friedman \cite{4} discussed the dynamical instability of neutral fluid sphere in the Newtonian as well as in the relativistic physics and showed that anisotropy enhances the stability if anisotropy is positive throughout matter distribution. Herrera et al. \cite{5} have studied the dynamical instability of the expansion free fluids using perturbation scheme. Different physical properties of the fluid plays important role in dynamical evaluation of the self-gravitating systems. According to Herrera et al. \cite{6,7} dissipation terms in the fluid would increase the instability of the collapsing objects. Chan and his collaborators \cite{8}-\cite{11} have showed that anisotropy and radiation would affect the instability range at Newtonian nd post-Newtonian approximation. Sharif and Azam \cite{11a}-\cite{14} have studied the effects of electromagnetic field on the dynamical stability of the collapsing dissipative and non dissipative fluids in spherical, cylindrical and plane symmetric geometries. This work has been further extended by Sharif and his collaborators \cite{15}-\cite{24} in higher order theories of gravity, like $f(R)$ and $f(T)$ and $f(R,T)$, in these papers the possible forms of the fluid with electromagnetic field have been discussed in detail. Till now many quantum theories of gravity have been proposed to investigate the natural phenomenon occurring in astronomy and astrophysics. Among these theories, superstring theory is the most strong candidate which has been extensively investigated for the spacetimes with more than four dimensions. In this theory the effects of extra dimensions becomes more prominent when curvature radius of the central high density regions during gravitational collapse becomes comparable with the curvature radius of the extra dimensions.From this point of view high density regions can be modeled in a sophisticated way in a theory which deals the extra dimensions. The braneworld universe model which is attractive proposal for the new picture of the universe is based on the superstring theory \cite{25}-\cite{30}. The geometrical interpretation of the braneworld model revels the fact that we are living on a four dimensional timelike hypersurface which is embedded in more than four dimensional manifold. This suggest that effects of superstring on the formation of back hole during the relativistic gravitational collapse of a star should be investigated explicitly. The current experiments performed for the tests of inverse square law do not exclude the possibility of the extra dimension even as large as a tenth of millimeter. As observed range of the gravitational force is directly dependent on the size of objects so it is interesting to consider some physical phenomena in the extra dimensions.On the basis of these facts it becomes important to study the general theory of relativity in more than four dimensions. In this regards a class of exact solutions to the Einstein field equations have been determined in the recent years \cite{31}-\cite{35}. These solutions play a significant role in studying the gravitational collapse evolution of the universe. Recently \cite{36}-\cite{43} there has been growing interest to study the higher order gravity , which are involves higher order derivatives of curvature terms. One of the most studied extensively higher order gravity theory is the Gauss-Bonnet gravity. This theory is the simplest generalization of general theory of relativity and special case of Lovelock Gravity theory. The Lagrangian of this theory contains just three terms as compared to Lagrangian of Loelock gravity Theory. The Gauss-Bonnet gravity theory is used to discuss the nontrivial dynamical systems in the dimensions greater or equal to 5. This theory naturally appears in the low energy effective action of the heterotic string theory. Boulware and Deser \cite{44} formulated the black hole (BH) solutions in N dimensional gravitational theory with four dimensional Gauss-Bonnet term. These are generalization of N dimensional solutions investigated by Tangherili \cite{45}, Merys and Perry \cite{46}. The spherically symmetric BH solutions and their physical properties have been studied in detail by Wheeller \cite{47}. The structure of topologically nontrivial BHs has been presented by Cai \cite{48}. Kobayashi \cite{49} and Maeda \cite{50} have explored the effects of Gauss-Bonnet term on the structure of Vaidya BH. All these studies show that appearance of the Gauss-Bonnet term in the field Equations would effect the occurrence of BH and Naked singularity during the gravitational collapse. In a recent paper \cite{52}, Jhinag and Ghosh have consider the $5D$ action with the Gauss-Bonnet terms in Tolman-Bondi model and give an exact model of the gravitational collapse of a inhomogeneous dust. Motivated by these studies, we have discussed the stability of the gravitationally collapsing spheres in Einstein Gauss-Bonnet gravity. This paper is organized as follow: In section \textbf{{2}} the Einstein Gauss-Bonnet field equations and dynamical equations have been presented. The perturbation scheme of first order on the field equations as well as on dynamical equations have been presented in section \textbf{3}. Section\textbf{ 4} deals with the Newtonian and post Newtonian approximation and derivation of the stability equation, which is main result of the paper. We summaries the results of the paper in the last section. \section{Interior Matter Distribution and Einstein Gauss-Bonnet Field Equations } We begin with the following 5D action: \begin{equation}\label{1} S=\int d^{5}x\sqrt{-g}\left[ \frac{1}{2k_{5}^{2}}\left( R+\alpha L_{GB}\right) \right] +S_{matter} \end{equation where $R$ ia a $5D$ Ricci scalar and $k^2_{5}={8\pi G_{5}}$ is $5D$ gravitational constant. The Gauss-Bonnet Lagrangian is of the form \begin{equation}\label{2} L_{GB}=R^{2}-4R_{ab}R^{ab}+R_{abcd}R^{abcd} \end{equation} where $\alpha$ is the coupling constant of the Gauss-Bonnet terms. This type of action is derived in the low-energy limit of heterotic superstring theory. In that case, $\alpha $ is regarded as the inverse string tension and positive definite and we consider only the case with $\alpha \geq 0$ in this paper. In the $4D$ space-time, the Gauss-Bonnet terms do not contribute to the Einstein field equations. The action (\ref{1}) leads to the following set of field equations \begin{equation}\label{3} {G}_{ab}=G_{ab}+\alpha H_{ab}=T_{ab}, \end{equation} where \begin{equation}\label{4} G_{ab}=R_{ab}-\frac{1}{2}g_{ab}R \end{equation} is the Einstein tensor and \begin{equation}\label{5} H_{ab}=2\left[ RR_{ab}-2R_{a\alpha }R_{b}^{\alpha }-2R^{\alpha \beta }R_{a\alpha b\beta }+R_{a}^{\alpha \beta \gamma }R_{b\alpha \beta \gamma \right] -\frac{1}{2}g_{ab}L_{GB}, \end{equation} is the Lanczos tensor. A spacelike 4D hypersurface $\Sigma^{(e)}$ is taken such that it divides a 5D spacetime into two 5D manifolds, $M^-$ and $M^+$, respectively. The 5D TB spacetime is taken as an interior manifold $M^-$ which represents an interior of a collapsing inhomogeneous and anisotropic sphere is given by \cite{52} \begin{equation}\label{6} ds_{-}^2=-dt^2+A^2dr^2+R^2(d\theta^2+\sin^2{\theta}d\phi^2 +\sin^2{\theta}\sin^2{\phi}d\psi^2), \end{equation} where $A$ and $R$ are functions of $t$ and $r$. The energy-momentum tensor $T_{\alpha \beta }^{-}$ for anisotropic fluid has the form \begin{equation}\label{7} T_{\alpha \beta }^{-}=(\mu +P_{\perp })V_{\alpha }V_{\beta }+P_{\perp }g_{\alpha \beta }+(P_{r}-P_{\perp })\chi _{\alpha }\chi _{\beta }, \end{equation} where $\mu $ is the energy density, $P_{r}$ the radial pressure, $P_{\perp }$the tangential pressure, $V^{\alpha }$ the four velocity of the fluid and $ \chi _{\alpha }$ a unit four vector along the radial direction. These quantities satisfy \begin{equation} V^{\alpha }V_{\alpha }=-1\ \ ,\ \ \ \ \ \chi ^{\alpha }\chi _{\alpha }=1\ \ ,\ \ \ \ \ \chi ^{\alpha }V_{\alpha }=0 \label{N8} \end{equation The expansion scalar $\Theta $ for the fluid is given by \begin{equation}\label{8} \Theta =V_{\ ;\ \alpha ,}^{\alpha }. \end{equation Since we assumed the metric (6) comoving, then \begin{equation}\label{9} V^{\alpha }=A^{-1}\delta _{0}^{\alpha }\ ,\ \ \ \ \ \chi ^{\alpha }=B^{-1}\delta _{1}^{\alpha }\ \end{equation and for the expansion scalar, we get \begin{equation}\label{10} \Theta =\frac{\dot{A}}{A}+\frac{3\dot{R}}{R}. \end{equation} Hence, Einstein Gauss-Bonnet field equations take the for \begin{eqnarray}\nonumber k^2_{5}\mu &&=\frac{12\left( R^{\prime 2}-A^{2}\left( 1+\dot{R}^{2}\right) \right) }{R^{3}A^{5}}\left[ R^{\prime }A^{\prime }+A^{2}\dot{R}\dot{A}-AR^{\prime \prime }\right] \alpha \\\label{11} &&\ \ \ \ \ -\frac{3}{A^{3}R^{2}}\left[ A^{3}\left( 1+\dot{R}^{2}\right) +A^{2}R\dot{R}\dot{A}+RR^{\prime }A^{\prime }-A(RR^{\prime \prime }+R^{\prime 2})\right]\\\label{12a} k^2_{5}p_{r} &&=-12\alpha \left( \frac{1}{R^{3}}-\frac{R^{^{\prime }2}}{A^{2}R^{3} +\frac{\dot{R}^{2}}{R^{3}}\right) \ddot{R}+3\frac{R^{^{\prime }2}}{A^{2}R^{2 } -3\Big(\frac{1+\dot{R}^{2}+R\ddot{R}}{R^{2}}\Big)\\\nonumber k^2_{5}p_{\perp } &&=\frac{4\alpha }{A^{4}R^{2}}\Big[ -2A\left( A^{^{\prime }}R^{^{\prime }}+A^{2}\dot{A}\dot{R}-AR^{^{\prime \prime }}\right) \ddot{R +A\left( R^{^{\prime }2}-A^{2}\left( 1+\dot{R}^{2}\right) \right) \ddot{A}\\\nonumber&&+2\Big( \dot{A}R^{^{\prime }}-A\dot{R}^{^{\prime }}\Big] -\frac{1}{A^{3}R^{2}}\Big[ A^{3}\Big( 1+\dot{R}^{2}+2R\ddot{R}\Big) +A^{2}R\left( 2\dot{R}\dot{A}+R\ddot{A}\right)\\&&+2RR^{^{\prime }}A^{^{\prime }}-2A\left( RR^{^{\prime \prime }}+R^{^{\prime }2}\right)\Big] \label{13}\\ &&\frac{12\alpha }{A^{5}R^{3}}\left( \dot{A}R^{^{\prime }}-A\dot{R}^{^{\prime }}\right) \left( A^{2}\left( 1+\dot{R}^{2}\right) -R^{^{\prime }2}\right) - \frac{A\dot{R}^{^{\prime }}-\dot{A}R^{^{\prime }}}{A^{3}R}=0 \label{14} \end{eqnarray} The mass function $m(t,r)$ analogous to Misner-Sharp mass in $n$ manifold without ${\Lambda}$ is given by \cite{50} \begin{equation}\label{15} m(t,r)=\frac{(n-2)}{2k_{n}^{2}}{V^k}_{n-2}\left[ R^{n-3}\left( k-g^{ab}R,_{a}R,_{b}\right) +(n-3)(n-4)\alpha \left( k-g^{ab}R,_{a}R,_{b}\right) ^{2} \right], \end{equation} where a comma denotes partial differentiation and ${V^k}_{n-2}$ is the surface of $(n-2)$ dimensional unit space. For $k=1$, ${V^1}_{n-2}=\frac{2{\pi}^{(n-1)/2}}{\Gamma((n-1)/2)}$, using this relation with $n=5$ and Eq.(\ref{6}), the mass function (\ref{15}) reduces to \begin{equation}\label{16} m(r,t)=\frac{3}{2}\left[ R^{2}\left( 1-\frac{R^{^{\prime }2}}{A^{2 }+\dot{R}^{2}\right) +2\alpha \left( 1-\frac{R^{^{\prime }2}}{A^{2}}+\dot{R ^{2}\right) ^{2}\right] \end{equation} The nontrivial components of the Binachi identities, $T_{;\beta }^{\ -\alpha \beta }=0$, from Eqs.(\ref{6}) and (\ref{7}), yield \begin{equation} \left[ \dot{\mu}+\left( \mu +P_{r}\right) \frac{\dot{A}}{A}+3\left( \mu +P_{\perp }\right) \frac{\dot{R}}{R}\right] =0 , \label{17} \end{equation an \begin{equation} T_{;\beta }^{\ -\alpha \beta }\chi _{\alpha }=\frac{1}{A}\left[ P_{r}^{^{\prime }}+3\left( P_{r}-P_{\perp }\right) \frac{R^{^{\prime }}}{R \right] =0 \label{18} \end{equation} Using field equations and Eq.(\ref{16}), we may write \begin{equation} m^{\prime }=\frac{2}{3}k^2_{5}\mu R^{\prime }R^{3} \label{N16a} \end{equation} In the exterior region to $\Sigma^{(e)}$, we consider Einstein Gauss-Bonnet Schwarzschild solution which is given by \cite{54} \begin{equation}\label{c1} ds_{+}^2=-F(\rho)d{\nu}^2-2d\nu d\rho+\rho^2(d\theta^2+\sin^2{\theta}d\phi^2 +\sin^2{\theta}\sin^2{\phi}d\psi^2), \end{equation} where $F(\rho)=1+\frac{{\rho}^2}{4\alpha}-\frac{{\rho}^2}{4\alpha}\sqrt{1+\frac{16\alpha M}{\pi {\rho}^4}}$ The smooth matching of the $5D$ anisotropic fluid sphere (\ref{6}) to GB Schwarzschild BH solution (\ref{c1}), across the interface at $r = {r_{\Sigma}}^{(e)}$ = constant, demands the continuity of the line elements and extrinsic curvature components (i.e., Darmois matching conditions), implying \begin{eqnarray}\label{c2} dt \overset{\Sigma^{(e)}}{=}\sqrt{F(\rho)}d\nu,\\ R \overset{\Sigma^{(e)}}{=}\rho, \\\label{cm} m(r,t)\overset{\Sigma^{(e)}}{=}M, \end{eqnarray} \begin{eqnarray}\nonumber &&-12\alpha \left( \frac{1}{R^{3}}-\frac{R^{^{\prime }2}}{A^{2}R^{3} +\frac{\dot{R}^{2}}{R^{3}}\right) \ddot{R}+3\frac{R^{^{\prime }2}}{A^{2}R^{2 } -3\Big(\frac{1+\dot{R}^{2}+R\ddot{R}}{R^{2}}\Big)\\ &&\overset{\Sigma^{(e)}}{=}\frac{12\alpha }{A^{5}R^{3}}\left( \dot{A}R^{^{\prime }}-A\dot{R}^{^{\prime }}\right) \left( A^{2}\left( 1+\dot{R}^{2}\right) -R^{^{\prime }2}\right) - \frac{A\dot{R}^{^{\prime }}-\dot{A}R^{^{\prime }}}{A^{3}R} \label{c3} \end{eqnarray} Comparing Eq.(\ref{c3}) with (\ref{12a}) and (\ref{14}) (for detail see \cite{12}), we get \begin{equation}\label{c4} p_r\overset{\Sigma^{(e)}}{=}0. \end{equation} Hence, the matching of the interior inhomogeneous anisotropic fluid sphere (\ref{6}) with the exterior vacuum Einstein Gauss-Bonnet spactime (\ref{c1}) produces Eqs.(\ref{6}) and (\ref{cm}). These are the necessary and sufficient conditions for the smooth matching of interior and exterior regions of a star on boundary surface ${\Sigma^{(e)}}$. It is well known that the expansionfree models present an internal vacuum cavity. The boundary surface between the external cavity and interior the fluid is labeled by ${\Sigma^{(i)}}$ then the smooth matching of the Minkowski spacetime within the cavity to the fluid distribution over ${\Sigma^{(i)}}$, yields \begin{eqnarray} m(r,t)\overset{\Sigma^{(i)}}{=}0.\\ p_r\overset{\Sigma^{(i)}}{=}0. \end{eqnarray} The physical applications of expansionfree models are wide in astrophysics and astronomy. For example, it may help to explore the structure of voids on cosmological scales\cite{55}. By definition Voids are the sponge like structures and occupying 40-50 percent of the entire universe. There are commonly two types of the voids: mini-voids \cite{56} and macro-voids\cite{57} On the basis of Observational data analysis the voids are neither empty nor spherical. For the sake of further exploration about voids they are considered as vacuum spherical cavities around the fluid distribution. \section{The Perturbation Scheme } In this section, we introduce the perturbation scheme, for this purpose it is assumed that initially fluid is in static equilibrium implying that the fluid is described by only such quantities which have only radial dependence. Such quantities are denoted by a subscript zero. We further assume as usual, that the metric functions $A(t,r)$ and $R(t,r)$ have the same time dependence in their perturbations. Therefore, we consider the metric and material functions in the following form \begin{eqnarray} A(t,r)&=&A_{0}(r)+\epsilon T(t)a(r), \label{N17}\\ R(t,r)&=&R_{0}(r)+\epsilon T(t)c(r), \label{N19}\\ \mu (t,r)&=&\mu _{0}(r)+\epsilon \bar{\mu}(t,r), \label{N20}\\ P_{r}(t,r)&=&P_{r0}(r)+\epsilon \bar{P}_{r}(t,r), \label{N21}\\ P_{\perp }(t,r)&=&P_{\perp 0}(r)+\epsilon \bar{P}_{\perp }(t,r), \label{N22}\\ m(t,r)&=&m_{0}(r)+\epsilon \bar{m}(t,r), \label{N23}\\ \Theta (t,r)&=&\epsilon \bar{\Theta}(t,r), \label{N24} \end{eqnarray} where $0<\epsilon \ll 1$ and we choose the Schwarzschild coordinates with $R_{0}(r)=r$. Using Eqs.(\ref{N17})-(\ref{N22}), we have from Eqs.(\ref{11})-(\ref{14}) the following static configuration \begin{eqnarray} k\mu _{0}&&=\frac{3}{r^{3}A_{0}^{4}}\left[ 4\alpha \left( \frac A_{0}^{^{\prime }}}{A_{0}}-A_{0}\right) -rA_{0}^{2}\left( A_{0}+\frac A_{0}^{^{\prime }}}{A_{0}}r-1\right) \right], \label{N25}\\ kP_{r0}&&=3\left[ \frac{1}{r^{2}A_{0}^{2}}-\frac{1}{r^{2}}-1\right], \label{N26}\\ kP_{\perp 0}&&=\frac{-1}{A_{0}^{2}r^{2}}\left[ A_{0}^{2}+2r\frac A_{0}^{^{\prime }}}{A_{0}}-2\right]. \label{N27} \end{eqnarray} Also from Eqs.(\ref{11})-(\ref{14}), we obtain the following form of the perturbed field equations \begin{eqnarray}\nonumber k\bar{\mu} &&=\frac{3T}{r^{2}A_{0}^{3}}\Big[ 4\alpha \Big( \frac{a^{\prime }}{rA_{0}^{2}}+\frac{3A_{0}^{\prime }c^{\prime }}{rA_{0}^{2}}-\frac c^{\prime \prime }}{rA_{0}}-\frac{a^{\prime }}{A_{0}r}-\frac{c^{\prime }A_{0}^{\prime }}{rA_{0}} \\\nonumber &&+ \frac{c^{\prime \prime }}{r}+\frac{3aA_{0}^{\prime }}{rA_{0}^{2}} \frac{5aA_{0}^{\prime }}{rA_{0}^{2}}-\frac{cA_{0}^{\prime }}{r^{2}A_{0}^{2}} \frac{cA_{0}^{\prime }}{r^{2}A_{0}}\Big) \\\nonumber &&-\Big( -3aA_{0}+2c-3ar\frac{A_{0}^{^{\prime }}}{A_{0}}+A_{0}^{\prime }c+ra^{\prime }\\\label{N28} && +A_{0}^{\prime }rc^{\prime }-A_{0}rc^{\prime \prime }-A_{0}+2a-2A_{0}c^{\prime }\Big) \Big] -\frac{2Tc}{r}k\mu _{0}\\\nonumber k\bar{P}_{r}&&=\frac{3\ddot{T}c}{r}\left[ 1-4\alpha \left( 1-\frac{1} r^{2}A_{0}^{2}}\right) \right] -\frac{6T}{r^{2}}\left( \frac{a}{A_{0}^{3}} \frac{c^{\prime }}{A_{0}^{2}}+cr\right) -\frac{2Tc}{r}kP_{r0},\\ \label{N29}\\\nonumber k^2_5\bar{P}_{\perp } &=&\frac{\ddot{T}}{A_{0}^{3}r^{2}}\left[ 4\alpha \left( a\left( 1-A_{0}^{2}\right) -2A_{0}^{^{\prime }}c\right) -A_{0}rc\left( A_{0}^{2}+r\right) \right] \\\nonumber &&+\frac{8\alpha \dot{T}}{A_{0}^{3}r^{2}}\left( \frac{a}{A_{0}}-c^{^{\prime }}\right) +\frac{T}{A_{0}^{2}r^{2}}\left[ 2rc^{^{\prime }}\left( \frac A_{0}^{^{\prime }}}{A_{0}}\right) +2a\left( \frac{A_{0}^{^{\prime }}}{A_{0} \right) -2rc^{^{\prime \prime }}\right. \\\label{N30} &&\left. +2r\left( \frac{a}{A_{0}}\right) ^{^{\prime }}-4c^{^{\prime }}r-5\left( \frac{a}{A_{0}}\right) -4ar\left( \frac{A_{0}^{^{\prime }}}{A_{0 }\right) \right] -\frac{2Tc}{r}kP_{\perp 0}, \\ &&\frac{12\alpha \dot{T}}{A_{0}^{5}r^{3}}\left( A_{0}^{2}a-A_{0}^{3}c-a-A_{0}c\right) -\frac{3\dot{T}}{A_{0}^{3}r}\left( A_{0}c^{\prime }-a\right) =0. \label{N31} \end{eqnarray} For the expansion given in Eq.(\ref{10}), we have \begin{equation} \bar{\Theta}=\dot{T}\left( \frac{a}{A_{0}}+\frac{3c}{R_{0}}\right). \label{N32} \end{equation The Binachi identities Eqs.(\ref{17}) and (\ref{18}) with (\ref{N17})-(\ref{N22}), yield the static configuration \begin{equation} P_{r0}^{\prime }+\frac{3}{r}\left( P_{r0}-P_{\perp 0}\right) =0 \label{N33} \end{equation and for the perturbed configuration \begin{eqnarray} \frac{1}{A_{0}}\left[ \bar{P}_{r}^{\prime }+\frac{3}{r}\left( \bar{P}_{r} \bar{P}_{\perp }\right) +3\left( P_{r0}-P_{\perp 0}\right) T\left( \frac{c}{ }\right) ^{\prime }\right] =0 , \label{N34}\\ \bar{\mu}=-\left[ \left( \mu _{0}+P_{r0}\right) \frac{a}{A_{0}}+\frac{3c}{r \left( \mu _{0}+P_{\perp 0}\right) \right] T . \label{N35} \end{eqnarray The total energy inside $\Sigma^{(e)\text{ }}$ up to a radius $r$ given by Eq.(\re {16}) with Eqs.(\ref{N17}),(\ref{N19}) and (\ref{N23}) becomes \begin{eqnarray} m_{0}&&=\frac{3}{2}\left[ \left( 1-\frac{1}{A_{0}^{2}}\right) \left( r^{2}+2\alpha \left( A_{0}^{2}-1\right) \right) \right], \label{N36}\\ \bar{m}&&=\frac{3T}{A_{0}^{2}}\left[ \left( A_{0}^{2}cr-c-c^{^{\prime }}r^{2}+r^{2}\frac{a}{A_{0}}\right) -\frac{\alpha }{A_{0}^{2}}\left( A_{0}^{2}-1\right) \left( c^{^{\prime }}-\frac{a}{A_{0}}\right) \right]. \label{N37} \end{eqnarray} From the matching condition Eq.(\ref{c4}), we have \begin{equation} P_{r0}\overset{\Sigma^{(e)}}{=}0,\ \ \ \ \bar{P}_{r}\overset{\Sigma^{(e)}}{=}0, \label{N38} \end{equation For $c\neq 0$, which is the case that we want to study , with (\ref{N29}), \ref{N31}) and (\ref{N38}) we obtain \begin{equation} \ddot{T}\ \beta -\gamma T=0, \label{N39} \end{equation} where \begin{equation*} \beta =1-4\alpha \left( 1-\frac{1}{r^{2}A_{0}^{2}}\right) ,\ \ \ \ \ \ \ \gamma =\frac{2}{rc}\left( \frac{a}{A_{0}^{3}}-\frac{c^{\prime }}{A_{0}^{2} +rc\right) \end{equation*} The general solution of Eq.(\ref{N39}) is actually the linear combination of two solutions one of these corresponding to stable (oscillating) system while other corresponds to unstable (non-oscillating) ones. As in the present case, we are interested to establish the range of instability, so we restrict our attention to the non oscillating ones, i.e., we assume that $a(r)$ and $c(r)$ attain such values on $r_{\Sigma ^{(e)}}$ that $\psi_{\Sigma ^{(e)}} =\Big(\frac{\beta }{\gamma }\Big) _{\Sigma ^{(e)}}>0.$ Then \begin{equation} T=\exp (-\sqrt{\psi _{\Sigma ^{(e)}}}t) \label{N40} \end{equation} representing collapsing sphere as areal radius becomes decreasing function of time. The dynamical instability of collapsing fluids can be well discussed in term of adiabatic index $\Gamma _{1}$. We relate$\bar{P}_{r}$ and $\bar{\mu}$ for the static spherically symmetric configuration as follows \begin{equation} \bar{P}_{r}=\Gamma _{1}\frac{P_{r0}}{\mu _{0}+P_{r0}}\bar{\mu}. \label{N41} \end{equation} We consider it constant throughout the fluid distribution or at least, throughout the region that we want to study. \section{Newtonian and Post Newtonian Terms and Dynamical Stability} This section deals to identify the Newtonian (N), post Newtonian (pN) and post post Newtonian (ppN) regimes. For this purpose we convert the relativistic units into c.g.s. units and expands all the terms in dynamical equations upto the $C^{-4}$ ($C$ being speed of light). In this analysis for the different regimes following approximation will be applicable \begin{itemize} \item N order: terms of order $C^0$; \item pN order: terms of order $C^{-2}$; \item ppN order: terms of order $C^{-4}$. \end{itemize} These terms are analyzed for the stability conditions appearing in the dynamical equation in the N approximations while pN and ppN are neglected. Thus, for N approximation, we assume \begin{equation} \mu _{0}\gg P_{r0},\ \ \ \ \ \mu _{0}\gg P_{\perp 0} \label{N42} \end{equation} For the metric coefficient expanded up to pN approximation, we take \begin{equation} A_{0}=1+\frac{Gm_{0}}{C^{2}r}, \label{N43} \end{equation} where $G$ is the gravitational constant and $C$ is the speed of light. With the help of equations obtained in previous sections, we can formulate the dynamical equation with expansion-free condition which is aim of our study. The key equation for the dynamical equation is Eq.(\ref{N34}). The expansion-free condition $\Theta =0$ implies from (\ref{N32}) \begin{equation} \frac{a}{A_{0}}=-3\frac{c}{r}, \label{N44} \end{equation} with (\ref{N44}), we have for (\ref{N35}) that \begin{equation} \bar{\mu}=3(P_{r0}-P_{\perp 0})T\frac{c}{r}. \label{N45} \end{equation} This equation explains how perturbed energy density of the system originates from the static background anisotropy. Also, with (\ref{N41}) and (\ref{N45}) we have \begin{equation} \bar{P}_{r}=3\Gamma _{1}\frac{P_{r0}}{\mu _{0}+P_{r0}}(P_{r0}-P_{\perp }) \frac{c}{r} \label{N46} \end{equation} From equations (\ref{N25})and (\ref{N36}), we have \begin{equation} \frac{A_{0}^{^{\prime }}}{A_{0}}=\frac{(r+m_{0})[(r+m_{0})^{3}k^2_5\mu _{0}+12\alpha ]}{12\alpha r-3r(r+m_{0})} \label{N47} \end{equation} Next, we develop dynamical equation by substituting Eq.(\ref{N30}) along with Eqs. (\ref{N44}),(\ref{N43}),(\ref{N39}) and (\ref{N47}) in Eq. (\re {N34}) and using the radial functions $a(r)=a_{0}r,\ c(r)=c_{0}r$, where $a_0$ and $c_0$ are constants. After a tedious algebra ( a detail procedure can be followed in \cite{6}), we obtain the dynamical equation at pN order ( with $c=G=1$) \begin{eqnarray}\nonumber &&\Big(12\alpha r-3r(r+m_{0})\Big)\Big[3\psi (r+m_{0})^2\Big(12\alpha c_{0}m_{0}(m_{0}+2r)-96\alpha ^{2}c_{0}r^{3}(r+m_{0}) \\\nonumber &&-c_{0}\Big((r+m_{0})^{2}+r^{3}\Big)\Big)+{8\alpha r^{8}\sqrt \psi }c_{0}}+3r^{3}c_{0}{(r+m_{0})}\Big(4r-15\Big) +6r^{3}(r+m_{0})^{3}c_{0}k^2_5P_{\perp 0} \\\nonumber &&+3r^{3}(r+m_{0})^{3}k^2_5(P_{r0}-P_{\perp 0})\Big]+{(r+m_{0})}\Big[216\alpha r^{3}c_{0}(1-2r)(r+m_{0})^{2}-72\alpha r^{4}c_{0}\Big] \\\nonumber &&=\Big[ 24r^3\alpha \psi c_{0}{(r+m_{0})}+{6r^4c_{0}} +18c_{0}r^3(2r-1){(r+m_{0})}\Big]k^2_5\lambda\Big(r^{n+1}+\frac{2}{3}(\frac{r^{n+4}}{n+4}-\frac{{r_i}^{n+4}}{n+4})\Big)\\\label{N48} \end{eqnarray} \begin{figure} \center\epsfig{file=1.eps, width=0.45\linewidth} \epsfig{file=c.eps, width=0.45\linewidth}\caption{The left graph is plotted for $\alpha=1,2,2.5$ and right graph is plotted for $c_0=-1,-3,-5$. For both graphs $ m_0=10$, $(P_{r0}-P_{\perp 0})=5, P_{\perp 0}=10$ are common} \center\epsfig{file=m.eps, width=0.45\linewidth}\epsfig{file=ptr.eps, width=0.45\linewidth}\caption{The left graph is plotted for $m_0=9.5, 10.5, 12$ and right graph is plotted for $(P_{r0}-P_{\perp 0})=2,4,6$. For both graphs $ \alpha=1$, $c_0=-2, P_{\perp 0}=10$ are common.} \epsfig{file=pr.eps, width=0.45\linewidth}\caption{This graph is plotted for $(P_{r0}-P_{\perp 0})=2$, $ \alpha=1$, $c_0=-2, P_{\perp 0}=10,12,13$.} \end{figure} \begin{figure} \center\epsfig{file=2a.eps, width=0.45\linewidth} \epsfig{file=2c.eps, width=0.45\linewidth}\caption{The left graph is plotted for $\alpha=1,2,2.5$ and right graph is plotted for $c_0=-1,-3,-5$. For both graphs $ m_0=10$, $\lambda=2, n=4$ are common} \center\epsfig{file=2m.eps, width=0.45\linewidth} \epsfig{file=2n.eps, width=0.45\linewidth}\caption{The left graph is plotted for $r_i=0.5, 0.7, 0.9$ and right graph is plotted for $n=2,4,6$. For both graphs $ \alpha=1$, $c_0=-2, \lambda=2, m_0=10$ are common.} \center\epsfig{file=L.eps, width=0.45\linewidth} \caption{ This graph is plotted for$ \alpha=1$, $c_0=-2, n=2, m_0=10, \lambda=2,4,6$ } \end{figure} Here, we have used Eq.(\ref{N16a}) and considered an energy density profile of the form $\mu _{0}=\lambda r^{n},$ where $\lambda $ is positive constant and $n$ is also a constant whose value ranges in the interval $-\infty <n<\infty .$ In order to fulfill the stability of expansion free fluids, we have to prove that both sides of Eq.(\ref{N48}) produces positive results, which is analytically impossible. We represent left side of Eq.(\ref{N48}) as $X(r)$ and right side of this equation by $Y(r)$. We prove graphically that for particulary values of the parameters involved in Eq.(\ref{N48}) both $X(r)$ and $Y(r)$ positive. The positivity of $X(r)$ and $Y(r)$ is shown in figures \textbf{(\textbf{1-3})} and \textbf{(\textbf{4-6})}, respectively. The values of the parameters for which $X(r)$ and $Y(r)$ remain positive (system predicts range of stability) are mentioned below each graph and other than these values system becomes unstable. \section{Summary} This paper deals with dynamical instability of the expansionfree anisotropic fluid at Newtonian and post Newtonian order in the frame work of Einstein Gauss-Bonnet gravity, which is vast play ground for higher dimensional analysis of general relativity . For a gravitating source which has non zero expansion scalar, the instability range of a self gravitating source can be defined by the adiabatic index $\Gamma_1$, which measures the compressibility of the fluid under consideration. On the other hand, for an expansionfree case as we are dealing, the instability explicitly depends upon the energy density, radial pressure, local anisotropy of pressure and Gauss-Bonnet coupling constant $\alpha$ at Newtonian approximation, but it appears to be independent of the adiabatic index $\Gamma_1$. In other words the stiffness of gravitating source at Newtonian and post Newtonian approximation does not play any role for the investigation of the stability of system. We would like to mention that anisotropy in pressure, inhomogeneity in the energy density and Gauss-Bonnet coupling constant $\alpha$ are the key factors for studying the the structure formation as well as evolution of shearfree anisotropic astrophysical objects. We have formulated two dynamical equations how gravitating objects evolve with time? and what is the final outcome of such evolution?. One of these dynamical equations is used to separate the terms which have Newtonian and post Newtonian order by using the concept of relativistic and c.g.s units. The post post Newtonian regimes are absent in the present analysis, it not due to Gauss-Bonnet gravity, it seems to occur due to the geodesic properties of the spacetime used in which $g_{00}=1$. This condition is in fact Newtonian limit of general relativity. The second dynamical equation is used to discuss the instability range of expansionfree fluid upto pN order. The first order perturbation scheme has been applied on the metric functions and matter variables appearing in the Gauss-Bonnet field equations and dynamical equations. The analysis of resulting dynamical equations shows that stability is independent of adiabatic index $\Gamma_1$ due to expansionfree fluid. The instability depends on the density profile, local anisotropy, Gauss-Bonnet coupling constant and some other parameters. The instability required that resultant of all term on left side of equation (\ref{N48}) should be positive and equal to resultant of all terms on right side of that equation. It is impossible to show analytically from Eq.(\ref{N48}), so we have proved this result for a particular values of the parameters appearing in Eq.(\ref{N48}). The domain of the parameters is taken conveniently to show both sides positive Fig. \textbf{(1-6)}. The parameters have following values for which system satisfies stability conditions: $1\leqslant\alpha2.5$, $-4\leqslant c_0\leqslant-1$,$9.5\leqslant m_0\leqslant12$, $2\leqslant(P_{r0}-P_{\perp 0})\leqslant6$, $10\leqslant P_{\perp 0} \leqslant13$, $0.5\leqslant r_i\leqslant0.9$, $2\leqslant\lambda15$, $4\leqslant n\leqslant 8.$ We have the novel values of the parameters one can carry actual calculations for the values of the parameters by introducing some restrictions on the system under consideration. This work with electromagnetic and heat flux in the presence of non-geodesic model i.e., $g_{00}\neq1$ is under progress \cite{58}. \vspace{0.25cm}
\section{Introduction} As a simple system, hydrogen atom with high precision measurements in atomic transitions is one of the best laboratories to test QED and new physics as well. Meanwhile, there is some discrepancy between the recent measurement of the muonic hydrogen Lamb shift and the corresponding proton radius and the CODATA value which is obtained from the spectroscopy of atomic hydrogen and electron-proton scattering \cite{codata}. There are two possibilities to explain the discrepancy: 1- the theoretical calculation within the standard model is incomplete. 2- existence of new physics beyond the standard model. There are attempts to explain the new physics by considering a new particle in MeV range where many stringent limits suppress its existence \cite{MeV}. However, the effective new interactions do not need necessarily new particles to mediate the new interactions \cite{newint}. For instance, noncommutative (NC) space can induce new interactions in QED without adding new particles in the theory. Theoretical aspects of the noncommutative space have been extensively studied by many physicists \cite{theo}. Meanwhile, noncommutative standard model (NCSM) via two different approaches is introduced in \cite{calmet,melic1,chaichian3} and its phenomenological aspects are explored in \cite{martin}. Here we would like to study two body bound state in noncommutative space to explore the differences in the electronic and muonic hydrogen spectrum. There are many studies on the hydrogen atom in the NC space-time \cite{ho,chaichian1,chaichian2,stern1,stern2,moumni,saha,bertolami,bertolami1,banerjee,Balachandran}. However, the effect of NC-space on the hydrogen atom at the lowest order is doubtful. P. M. Ho and H. C. Kao \cite{ho} have shown that there is not any correction on the space-space NC-parameter in this system. Since all fields live on the same noncommutative space the noncommutativity of a particle not only should be opposite to its anti-particle but also the NC-parameter of a charged particle should be opposite to any other particle of opposite charge. In fact, for the proton as a point particle $\theta_p=-\theta_e$ and the corrections on the Coulomb potential coming from both particles cancel out each other. However, the proton is not a point particle and the parameter of noncommutativity is an effective parameter and is not equal to $-\theta_e$ \cite{chaichian2}. Furthermore, even if the proton can be considered as a point particle, the space time noncommutativity has some impact on the spectrum of the atom. Meanwhile, in the Lorentz conserving NCQED the NC-parameter appears as $\theta^2$ which is equal for a point particle and its antiparticle. Therefore, it is reasonable to examine the hydrogen atom in the NC space. As $\theta_{\mu\nu}$ has dimension $-2$ one expects an energy shift proportional to $(\theta m^2)^2mc^2\alpha^2$. In fact, in the NC-space a larger energy shift for the muonic Hydrogen is expected in comparison with the ordinary Hydrogen which is in agreement with the experimental data. In Section II we explore two-body bound state in NC-space. In section III we examine the $1S$-$2S$ transition and the Lamb-shift for Hydrogen and muonic-Hydrogen and the $g$-factor for electron and muon. In section IV we give a brief review on the Lorentz conserving NCQED and calculate the $g$-factor for electron and muon, the $1S$-$2S$ transition and the Lamb-shift for Hydrogen and muonic-Hydrogen. In section V we summarize our results. \section{2-body bound state in Noncommutative Space-Time} Since the NC-parameters for a point particle and its antiparticle are opposite, the NC-correction on the potential for the space-space part of NC-parameter is zero at the lowest order. It can be shown that the Coulomb potential in the Schrodinger equation is proportional to $(\theta_p+\theta_e)_{ij}$ that is zero in the point particle limit of the proton \cite{Stefano, ho}. In the both references the starting point is the Schrodinger equation in the NC space. However, one can show how the NC field theory through the Bethe-Salpeter (BS)-equation leads to the Schrodinger equation with a modified potential\cite{2-body}. In fact, it is better to start from the NC-field theory (NCFT) to avoid the mistake on the NC-contributions from each particle in the bound state. For instance, the correct potential in the NC-Schrodinger equation can be explored in studying the kernel in the BS-equation for the corresponding NCFT. For this purpose examining the electron-proton scattering amplitude in the NCQED is adequate to derive the appropriate potential for the NCQM. In a canonical noncommutative space, space-time coordinates are not numbers but operators which do not commute \begin{equation}\label{103} [\hat{x}^{\mu},\hat{x}^{\nu}]=i\theta^{\mu\nu}=i\frac{C^{\mu\nu}}{\Lambda_{\tiny{NC}}^2}, \end{equation} where $\theta^{\mu\nu}$ is the parameter of noncommutativity, $C^{\mu\nu}$ is a constant and dimensionless antisymmetric tensor and $\Lambda_{NC}$ is the noncommutative scale. Since noncommutative parameter, $\theta^{\mu\nu}$, is constant and identifies a preferred direction in space, canonical version of non-commutative space-time leads to the Lorentz symmetry violation. There are two versions to construct the NCQED \cite{calmet,melic1,chaichian3}. In the first one in contrast with the ordinary QED a momentum dependent phase factor appears in the charged fermion-photon vertex as follows \cite{riad} \begin{equation}\label{1} ieQ\gamma_\mu \exp{(ip_\mu\theta^{\mu\nu}p'_\nu)}, \end{equation} where $p_\mu$ and $p'_\nu$ are the incoming and outgoing momenta and $\theta$ is the NC-parameter. Therefore, the electron-proton amplitude can be written as \begin{equation}\label{2} {\cal{M}}_{NC}={\cal{M}}\exp({ip_\mu(\theta_e)^{\mu\nu}p'_\nu})\exp({ik_\mu(\theta_p)^{\mu\nu}k'_\nu}), \end{equation} where $p$ ($p'$) and $k$ ($k'$) are the incoming (outgoing) momenta of the electron and proton, respectively. One can easily see that the exponent in terms of the momentum transfer $q$ in the center of mass is \begin{equation}\label{3} -i(\vec{\theta}_e+\vec{\theta}_p).\vec{p}\times\vec{q} -ip_i(\theta_e^{i0}+\theta_p^{i0})q_0-i(p_0\theta_e^{0i}-k_0\theta_p^{0i})q_i, \end{equation} where for $\theta_p=-\theta_e=\theta$ results in \begin{equation}\label{4} i(\sqrt{m_e^2+\bf{p}^2}+\sqrt{m_p^2+\bf{p}^2})\theta^{0i}q_i. \end{equation} Therefore, for $m^2_e$ and $m^2_p\gg\bf{p}^2$, the non-relativistic potential in the NCQM is the Fourier transform of \begin{equation}\label{5} \frac{e^2\exp[i(m_e+m_p)\theta^{0i}q_i]}{q^2}, \end{equation} that leads to $V_{NC}=V(\vec{r}-(m_e+m_p)\vec{\theta}_t)$ where $\vec{\theta}_t=(\theta_{10}, \theta_{20}, \theta_{30})$. In fact, to the lowest order of both $\theta$ and $\alpha$ the hydrogen atom only receives some contribution from the temporal part of the NC-parameter. Meanwhile, in \cite{stern1} and \cite{stern2} a new correction due to the non commutativity of the source at the lowest order of the space part of the NC-parameter is found. Nonetheless, it is of the order $\alpha^6$ and is not at the lowest order of $\alpha$ too. In the second approach, via Seiberg-Witten maps, the fields also depend on the noncommutative parameter. Using Seiberg-Witten maps, noncommutative standard model and Feynman rules are fully provided in references \cite{calmet,melic1}. In this approach, two fermion-photon vertex is \cite{melic1} \begin{equation}\label{104} ieQ_{f}\gamma_{\mu}+\frac{1}{2}eQ_{f}[(p_{out}\theta p_{in})\gamma_{\mu}-(p_{out}\theta)({\p}_{in}-m_{f})-({\p}_{out}-m_{f})(\theta p_{in})_{\mu}] , \end{equation} Where $Q$ is the fermion charge and $p_{in}$ and $p_{out}$ are incoming and outgoing momenta, respectively. Considering proton as a point particle, scattering amplitude of electron-proton in the on-shell limit can be given as follows \begin{equation}\label{105} iM=\bar{u}(p')[-ie\gamma^{\mu}-\frac{1}{2}e(p'\theta_{e} p)\gamma^{\mu}]u(p)(- ig_{\mu\nu}/ q^{2})\bar{\nu}(k)[ie\gamma^{\mu}+\frac{1}{2}e(k'\theta_{p} k)\gamma^{\mu}]\nu(k'), \end{equation} where $p$ ($k$) and $p'$ ($k'$) are incoming and outgoing momenta of electron (proton), respectively. At the lowest order of $\theta_{\mu\nu}$, (\ref{105}) is the same as (\ref{2}) that means in the second approach one has the same result as is given in (\ref{5}). In fact, for point particles in the QED bound states such as positronium, there is no NC-correction at the lowest order of $\alpha$ and $\theta_{ij}$. \section{Hydrogen Atom in Noncommutative Space-Time } In a two body bound state the point particles satisfy $\theta_p=-\theta_e$ and the interaction potential only depends on the time part of the NC-parameter as is shown in (\ref{5}). Nevertheless, proton in the Hydrogen atom is not a point particle and has an effective NC-parameter in terms of the NC-parameters of its contents \cite{chaichian2} which is not equal to $-\theta_e$. In fact, for the NC-parameter of the order of $1 TeV$ either the electron with energy of order $eV$ or muon with $keV$-energy in the Hydrogen cannot probe inside the proton to see its noncommutativity. It should be noted that for the muon the Bohr radius is about $\frac{m_e}{m_{\mu}}a_e\sim 2.5 \times 10^{-13} m\sim 3r_p$ which is comparable with the size of the proton. Meanwhile, for instance in the Lamb shift of muon-Hydrogen the hadronic interactions are of the order of $\alpha^2$ smaller than the QED interaction of the muon with proton as a point particle. Thus, considering the NC-effects on the muon interaction with the proton contents leads to corrections of the order $m_{\mu}^2\theta\alpha^2$ smaller than what one finds in the main part of the interactions. Therefore, even in the muon-Hydrogen the NC-corrections due to the finite size of the proton is negligible. In fact, at the low energy limit from the noncommutative point of view proton is a macroscopic particle and $\theta_{proton}\simeq 0$. Therefore, in the non-relativistic limit and for the energy scale of atom, equation (\ref{3}) leads to \begin{equation}\label{33} -i\vec{\theta}_e.\vec{p}\times\vec{q} -ip_i\theta_e^{i0}q_0-ip_0\theta_e^{0i}q_i. \end{equation} Equation (\ref{33}) for $\theta^{0i}= 0$ and $\theta^{ij}\neq 0$, has been already considered to find bound on the NC-parameter in Hydrogen atom \cite{chaichian1,stern1,stern2,moumni}. The temporal part of noncommutativity has some problem with the unitarity. However, in some cases it can be shown that the quantum mechanics is unitary for the temporal part of NC-parameter \cite{Balachandran1,Balachandran2}. Nevertheless, we examine the temporal part ($\theta^{0i}\neq 0$ and $\theta^{ij}= 0$) in the non-relativistic limit where (\ref{33}) leads to \begin{equation}\label{poten1} \frac{e^2\exp[im_e\theta^{0i}q_i]}{q^2}, \end{equation} or \begin{equation}\label{1061} V_{nc}(\vec{r})=V(\vec{r}-m_e\vec{\theta}_t)-V(\vec{r})\simeq - \alpha m_{e}\frac{\overrightarrow{\theta}.\overrightarrow{r}}{r^{3}}. \end{equation} In (\ref{1061}) only the parallel part of $\vec{\theta}$ with $\vec{r}$ has some contribution on the NC-potential that is \begin{equation}\label{1071} V_{nc}(r)= -m_{e} \alpha \frac{|\vec{\theta}|\cos{\phi}}{r^{2}}, \end{equation} where $\cos{\phi}$ is the angle between $\vec{\theta}$ and $\vec{r}$ and $\alpha $ is the fine structure constant. Using the perturbation theory for the ground state leads to zero energy shift for this state. However, for the excited states in the non-relativistic limit, the states $^2S_{\frac{1}{2}}$ and $^2P_{\frac{1}{2}}$ are degenerate. Therefore, the potential (\ref{1071}) splits these states into 2 states with an energy difference proportional to $<R_{n0}\mid\frac{-m_e\alpha |\vec{\theta}|}{r^2}\mid R_{n1}>\sim -m^3_e\alpha^3 |\vec{\theta}|$. But one can surprisingly show that \begin{equation}\label{108'} <R_{20}\mid\frac{1}{r^2}\mid R_{21}>=0, \end{equation} or up to the first order of $|\vec{\theta}|$ the states $2^2S_{\frac{1}{2}}$ and $2^2P_{\frac{1}{2}}$ remain degenerate. Nonetheless, for $n=3$ one has \begin{equation}\label{108} \Delta E_{NC}^{H-atom}=\frac{2\sqrt{2}}{3}m^3_e\alpha^3 |\vec{\theta}|. \end{equation} Here we consider the effects of the NC-space-time on the physical quantities such as $1S-3S$ transition, Lamb-shift in atom and the anomalous magnetic moment to fix the NC-parameter in accordance with the experimental uncertainties. {\bf $1S-3S$ Transition:} The obtained energy shift (\ref{108}), from the temporal part of noncommutativity, leads to an additional contribution on the theoretical value of $1S$-$3S$ transition in Hydrogen atom as follows \begin{equation}\label{108'} \Delta E_{NC}^{1S-3S}=\frac{2\sqrt{2}}{3}m^3_e\alpha^3 |\vec{\theta}|. \end{equation} The uncertainty on the experimental value for $1S$-$3S$ transition in Hydrogen atom is about $13\hspace{.2cm} kHz $\cite{1s3s} \begin{equation}\label{109} f_{1S-3S}=2 922 742 936.729 (13) MHz. \end{equation} Therefore, for $\Lambda=1.5\hspace{.2cm}TeV$ one has \begin{equation}\label{110} \Delta E_{NC}^{1S-3S}\sim 30\hspace{.2cm} Hz\ll 13\hspace{.2cm} kHz. \end{equation} {\bf Lamb Shift:} For $\Lambda=1.5\hspace{.2cm}TeV$ the NC-correction on the lamb shift for the electronic hydrogen is \begin{equation}\label{lamb-e} \Delta E_{NC}^{H_e}=\frac{2\sqrt{2}}{3}m^3_e\alpha^3 |\vec{\theta}|\sim 30\hspace{.2cm} Hz\ll 48\hspace{.2cm} kHz, \end{equation} where $48\hspace{.2cm} kHz$ is the experimental accuracy on the $n=3$ lamb shifts in the Hydrogen atom\cite{lamb3s}. Meanwhile, for the muonic hydrogen one has \begin{equation}\label{lamb-mu} \Delta E_{NC}^{H_{\mu}}=(\frac{m_{\mu}}{m_e})^{3}\Delta E_{NC}^{H_e}= 240\hspace{.2cm} MHz\sim 2\times10^{-4} meV. \end{equation} {\bf g-2 for electron and muon:} Since the NC-parameter has dimension -2 the dimensionless quantity $a=\frac{g-2}{2}$ should be corrected, at the lowest order in NC-space, as $C(\frac{\alpha}{2\pi})p_\mu\theta^{\mu\nu}p'_\nu$, where $C$ is a constant which is obtained in \cite{mortazavi,riad}. Therefore, for $\Lambda=1.5\hspace{.2cm} TeV$ the NC-correction on $a$ for electron is \begin{equation}\label{lamb-e} \delta a_e=\frac{5}{6}\frac{\alpha}{2\pi}\frac{p^2}{\Lambda^2} \simeq\frac{5}{6}\frac{\alpha}{2\pi}\frac{m_e^2}{\Lambda^2}\sim10^{-16}, \end{equation} and for the muon where in E286 experiment has a momentum about $3\hspace{.2cm} GeV$ is \begin{equation}\label{lamb-e} \delta a_\mu=\frac{5}{6}\frac{\alpha}{2\pi}\frac{p^2}{\Lambda^2} \simeq\frac{5}{6}\frac{\alpha}{2\pi}(\frac{3}{1500})^2\sim 3\times10^{-9}. \end{equation} \section{Hydrogen atom in Lorentz Conserving Noncommutative Space-Time} As the NC-parameter is a real and constant Lorentz tensor, there is, obviously, a preferred direction in a given particle Lorentz frame which leads to the Lorentz symmetry violation. On the other hand, experimental inspections for Lorentz violation, including clock comparison tests, polarization measurements on the light from distant galaxies, analyses of the radiation emitted by energetic astrophysical sources, studies of matter-antimatter asymmetries for trapped charged particles and bound state systems \cite{LV} and so on, have thus far failed to produce any positive results. These experiments strictly bound the Lorentz-violating parameters, therefore, in the lower energy limit, the Lorentz symmetry is an almost exact symmetry of the nature \cite{wolf}. However, Carlson, Carone, and Zobin (CCZ) have constructed Lorentz-conserving noncommutative quantum electrodynamics based on a contracted Snyder algebra \cite{Carlson}. In this class of NC theories, the parameter of noncommutativity is not a constant but an operator which transforms as a Lorentz tensor. In fact, (\ref{103}) should be extended to \begin{eqnarray}\label{103'} [\hat{x}^{\mu},\hat{x}^{\nu}]=i\hat{\theta}^{\mu\nu},\,\,\, [\hat{\theta}^{\alpha\beta},\hat{\theta}^{\mu\nu}]=0,\,\,\,[\hat{\theta}^{\mu\nu},\hat{x}^{\nu}]=0, \end{eqnarray} where $\hat{\theta}^{\mu\nu}$ is an operator. Consequently, according to the Weyl- Moyal correspondence, to construct the LCNC action the ordinary product should be replaced with the star product as follows \begin{equation}\label{starpro} f\ast g(x,\hat{\theta)}= f(x,\theta)exp(i/2\overleftarrow{\partial}_{\mu}\theta^{\mu\nu}\overrightarrow{\partial}_{\nu})g(x,\theta). \end{equation} In this formalism a sufficiently fast falling weight function $W(\theta)$ has been used to construct the Lorentz invariant lagrangian in a non-commutative space as \begin{equation}\label{110'} \mathcal{L}(x)=\int d^{6}\theta W(\theta) \mathcal{L}(\phi,\partial\phi)_{\ast}, \end{equation} where the Lorentz invariant weight function $W(\theta)$ is introduced to suppresses the NC-cross section for energies beyond the NC-energy scale. In reference \cite{exist} on the existence of an invariant normalized weight function is discussed and an explicit form for $W(\theta)$ is given in terms of Lorentz invariant combinations of $\theta^{\mu\nu}$'s. The function $W(\theta)$ can be used to define an operator trace as \begin{equation}\label{106} Tr\hat{f}=\int d^{4}x d^{6}\theta W(\theta)f(x,\theta), \end{equation} in which $W(\theta)$ has the following properties \begin{equation} \int d^{6}\theta W(\theta) = 1, \end{equation} \begin{equation}\label{107} \int d^{6}\theta W(\theta)\theta^{\mu\nu}=0, \end{equation} \begin{equation}\label{1082} \int d^{6}\theta W(\theta)\theta^{\mu\nu}\theta^{\kappa\lambda}= \langle\frac{\theta ^{2}}{2}\rangle (g^{\mu\nu}g^{\mu\lambda}-g^{\mu\lambda}g^{\nu\kappa}), \end{equation} where \begin{equation}\label{109'} \langle\theta^{2}\rangle=\int d^{6}\theta W(\theta)\theta^{\mu\nu}\theta_{\mu\nu}. \end{equation} As (\ref{107}) shows in the expansion of the Lagrangian (\ref{110'}) in terms of the NC-parameter the odd powers of $\theta_{\mu\nu}$ vanishes. In fact, to obtain the nonvanishing $\theta$-dependence terms, all fields should be expanded at least up to the second order of the NC-parameter. The Lorentz conserving NCSM is fully introduced in \cite{haghighat} and its fermionic part where we are interested to explore is given as follows \begin{eqnarray}\label{111} S_{fermion}= \int d^{4}x (\overline{L}i{\D}L + \overline{R}i{\D}R)+ \int d^{6}\theta \int d^{4}x W(\theta)\theta^{\mu\nu}\theta^{\kappa\lambda} (-\frac{i}{8}\overline{L} \gamma^{\rho}F^{0}_{\mu\kappa}F^{0}_{\lambda\rho}D^{0}_{\nu}L\no\\ -\frac{i}{4}\overline{L}\gamma^{\rho}F^{0}_{\mu\rho}F^{0}_{\nu\kappa}D^{0}_{\lambda}L -\frac{1}{8}\overline{L}\gamma^{\rho}(D^{0}_{\mu}F^{0}_{\kappa\rho})D^{0}_{\nu}D^{0}_{\lambda}L -\frac{i}{8}\overline{L} \gamma^{\rho}F^{0}_{\mu\nu}F^{0}_{\kappa\rho}D^{0}_{\lambda}L)\hspace{1cm} \no\\ \int d^{6}\theta \int d^{4}x W(\theta)\theta^{\mu\nu}\theta^{\kappa\lambda} (-\frac{i}{8}\overline{R} \gamma^{\rho}F^{0}_{\mu\kappa}F^{0}_{\lambda\rho}D^{0}_{\nu}R -\frac{i}{4}\overline{R}\gamma^{\rho}F^{0}_{\mu\rho}F^{0}_{\nu\kappa}D^{0}_{\lambda}R\hspace{1cm}\no\\ -\frac{1}{8}\overline{R}\gamma^{\rho}(D^{0}_{\mu}F^{0}_{\kappa\rho})D^{0}_{\nu}D^{0}_{\lambda}R -\frac{i}{8}\overline{R} \gamma^{\rho}F^{0}_{\mu\nu}F^{0}_{\kappa\rho}D^{0}_{\lambda}R),\hspace{3cm} \end{eqnarray} where $L$ and $R$ stand, respectively, for left and right handed fermions and $F^{0}_{\mu\nu}$ is the ordinary field strength in the standard model. To find the LCNC-effects, at the lowest order, on the hydrogen atom we only consider the QED part of the NC-action (\ref{111}) as follows \begin{eqnarray}\label{112} \int d^{6}\theta \int d^{4}x eQ_{f}(\overline{\nu}_{L}{\A}_{0}\nu_{L}-\frac{1}{8}\theta^{\mu\nu}\theta^{\kappa\lambda}\overline{\nu}_{L}\gamma^{\rho} \partial_{\mu}A_{0\kappa\rho}\partial_{\nu}\partial_{\lambda}\nu_{L}+\overline{e}{\A}_{0}e_{L}-\no\\ \frac{1}{8}\theta^{\mu\nu}\theta^{\kappa\lambda}\overline{e}\gamma^{\rho}\partial_{\mu}A_{0\kappa\rho}\partial_{\nu}\partial_{\lambda}e),\hspace{3cm} \end{eqnarray} Where $A_{0\mu\nu} = \partial_{\mu}A_{0\nu}-\partial_{\nu}A_{0\mu}$ and the charged fermions interact with photon via the following vertex \begin{equation}\label{113} ieQ_{f}\gamma^{\mu}(1+ \frac{\langle\theta^{2}\rangle}{96}(\frac{q^{4}}{4}-m_{f}^{2}q^{2}) ). \end{equation} In (\ref{113}) the $\theta$-dependence is appeared as $\langle\theta^{2}\rangle$ which is similar for both particle and its antiparticle. In fact, in LCNC-QED in contrast with NCQED according to $\theta_{f_-}= -\theta_{f_+}$, particle vertex doesn't cancel antiparticle vertex. Therefore, in $f_-f_+$ bound state the LCNC effect via the $f_-f_+$ scattering amplitude, at low energy limit, leads to a potential in momentum space as \begin{equation}\label{114} \widetilde{V}(q)= -\frac{e^{2}}{q^{2}}-\frac{e^{2}(m_{f^-}^{2}+m_{f^+}^{2})\langle \theta^{2}\rangle}{96}, \end{equation} Where to obtain (\ref{114}), at low momentum transfer, the second term in (\ref{113}) is ignored in comparison with the third one. However, proton as a particle with internal structure doesn't see the NC-space in those systems which the momentum transfer is small such as Hydrogen like atom. Therefore, the NC-potential in the Hydrogen atom is \begin{equation}\label{hyd} \widetilde{V}(q)= -\frac{e^{2}}{q^{2}}-\frac{e^{2}m_{e^-}^{2}\langle \theta^{2}\rangle}{96}, \end{equation} or \begin{equation}\label{115} V(r) =- \frac{e^{2}}{4\pi r}-\frac{e^{2}m_{e}^{2}\langle \theta^{2}\rangle}{96}\delta(r). \end{equation} The NC-correction on the Coulomb potential in (\ref{115}) is small and its expectation value directly gives the energy shift on the energy levels of the Hydrogen atom as follows \begin{equation}\label{117} \Delta E_{LCNC}^{H-atom}= - \langle \psi|\frac{e^{2}m_{e^-}^{2}\langle \theta^{2}\rangle}{96}\delta(r)|\psi\rangle= -\frac{e^{2}m_{e^-}^{2}\langle \theta^{2}\rangle}{96} |\psi_{nl}(r=0)|^{2} , \end{equation} Where $ |\psi_{nl}(r=0)|^{2}=\frac{\alpha^{3}m_{e^-}^{3}}{\pi n^{3}}\delta_{l0}$ leads to \begin{equation}\label{118} \Delta E_{LCNC}^{H-atom}= - \frac{m_{e^-}^{5}\alpha^{4}\langle\theta^{2}\rangle}{24n^{3}} \delta_{l0}, \end{equation} or with $\Lambda_{LCNC}=(\frac{12}{\langle\theta^{2}\rangle})^{4}$, (\ref{118}) can be rewritten as \begin{equation}\label{119} \Delta E_{LCNC}^{H-atom}= -\frac{m_{e^-}^{5}\alpha^{4}}{2n^{3}}\frac{1}{\Lambda_{LCNC}^{4}}\delta_{l0}. \end{equation} Here we fix the NC-parameter by the most precise experimental value in the Hydrogen atom (i.e. $1S-2S$ transition) then we find the effects of the NC-space-time on the other physical quantities such as Lamb-shift in atom and the anomalous magnetic moment. {\bf $1S-2S$ Transition:} The experimental value for $1S$-$2S$ transition in the Hydrogen atom \cite{phil} can fix the upper bound on the parameter $\Lambda_{LCNC}$ as follows \begin{equation}\label{1s2s} \Delta E_{LCNC}^{1S-2S}= \frac{7m_{e}^{5}\alpha^{4}}{16}\frac{1}{\Lambda_{LCNC}^{4}}\sim 34\hspace{.2cm} Hz, \end{equation} leads to \begin{equation}\label{119} \Lambda_{LCNC}\sim 0.2\hspace{.2cm} GeV. \end{equation} {\bf Lamb Shift:} For $\Lambda=0.5\hspace{.2cm} GeV$ the NC-correction on the lamb shift for electronic hydrogen is \begin{equation}\label{lc-lamb-e} \Delta E_{NC}^{H_e}=\frac{m_{e}^{5}\alpha^{4}}{16}\frac{1}{\Lambda_{LCNC}^{4}}\sim 0.1\hspace{.2cm} Hz \ll 3\hspace{.2cm} kHz, \end{equation} where $48\hspace{.2cm} kHz$ is the experimental accuracy on the $n=3$ lamb shifts in the Hydrogen atom\cite{eides}. Meanwhile, for the muonic hydrogen one has \begin{equation}\label{lc-lamb-mu} \Delta E_{NC}^{H_{\mu}}=(\frac{m_{\mu}}{m_e})^{5}\Delta E_{NC}^{H_e}\simeq 40 \hspace{.2cm}GHz\sim 0.03\hspace{.2cm} meV. \end{equation} {\bf g-2 for electron and muon:} As Eq.(\ref{113}) shows the NC-correction on $a=\frac{g-2}{2}$ should be proportional to $q^2$ which leads to zero NC-correction on $a$ at zero momentum transfer. \section{summary} In this paper two body bound state has been studied by examining the scattering amplitude in the Lorenz violated (LV) and Lorentz conserving (LC) NCQED as given in (\ref{2}) and (\ref{114}), respectively. For a bound state of a particle with its antiparticle the NC potential in LVNCQED, in contrast with LCNCQED, only depends on the space-time part of the NC-parameter, see (\ref{3}). As the proton in Hydrogen atom is not a point particle, $\theta_p\neq-\theta_e$. In fact, in $ep$-scattering in the low energy limit, the electron cannot prob inside the proton to see its NC-effects. In the Lorentz violated NCQED for Hydrogen and muonic Hydrogen atom we have found:\\ 1- In $1S-2S$ transition in the Hydrogen atom the NC-effect is zero, see(\ref{108'}).\\ 2- In $1S-3S$ transition in the Hydrogen atom the NC-parameter of the order of $\Lambda_{NC}=1.5\,\, TeV$ leads to a small correction on the theoretical value which is not detectable.\\ 3- $\Lambda_{NC}=1.5\,\, TeV$ leads to an NC-shift on the $n=3$ Lamb shift about $30\,\,Hz$ which is far from the $48\,\,kHz$ current uncertainty on the Lamb shift in the Hydrogen atom. As (\ref{lamb-mu}) shows an uncertainty of order of $3\,\,kHz$ in the Lamb shift of Hydrogen leads to $(3\,kHz)(\frac{m_\mu}{m_e})^3=26\,\,GHz$ which can only explain a small part of the current deviation between the experimental measurement and the theoretical prediction. \\ 4- The NC-effect on the g factors of electron and muon are $a_e=10^{-16} $ and $a_\mu=10^{-9}$, respectively. The obtained values for $a_e$ and $a_\mu$ are in agreement with the experimental measurements.\\ 5-$\Lambda_{NC}=1.5\,\, TeV$ which is found from the muon g-factor is a stringent bound in the atomic systems where the NC-parameter is usually of the order of a few $GeV$ \cite{stern1,stern2}. However, this bound can not explain the discrepancy in measurement of Lamb shift in muonic Hydrogen.\\ In the Lorentz conserving NCQED for Hydrogen and muonic Hydrogen atom we have found:\\ 1- In $1S-2S$ transition in the Hydrogen atom the NC-parameter has been fixed about $\Lambda_{NC}=0.5\,\, GeV$ which is the first bound on this parameter in an atomic system.\\ 2- $\Lambda_{NC}=0.5\,\, GeV$ leads to an NC-shift on the $2s_{1/2}-2p_{1/2}$ transition in hydrogen atom about $0.1\,\,Hz$ which is far from the $3\,\,kHz$ current uncertainty on the Lamb shift in the Hydrogen atom. As (\ref{lc-lamb-mu}) shows an uncertainty of order of $3\,\,Hz\ll 3\,\,kHz$ in the Lamb shift of Hydrogen leads to $(3\,Hz)(\frac{m_\mu}{m_e})^5=1000\,\,GHz$ which can explain the current difference between the experimental measurement and the theoretical prediction about $0.3\, meV$ \cite{codata}. \\ 3- The NC-correction on $a=\frac{g-2}{2}$ is proportional to $q^2$ which is zero at the zero momentum transfer. \\
\section{acknowledgements} This work is part of the research program of the Foundation for Fundamental Research on Matter (FOM), which is financially supported by the Netherlands Organisation for Scientific Research (NWO). S.Y.T.v.d.M. acknowledges support from NWO via a VIDI and a TOP grant, and from the European Research Council via a Starting Grant. We thank Sean Gordon for assistance during a research visit and David Parker for fruitful discussions. The expert technical support by Leander Gerritsen, Chris Berkhout, Peter Claus, Niek Janssen and Andr\'e van Roij is gratefully acknowledged.
\section{Introduction} In many statistical learning problems, the energy functions to be optimized are highly non-convex. A large body of research has been devoted to either approximating the target function by convex optimization, such as replacing $L_0$ norm by $L_1$ norm in regression, or designing algorithms to find a good local optimum, such as EM algorithm for clustering. Much less work has been done in analyzing the properties of such non-convex energy landscapes. In this paper, inspired by the success of visualizing the landscapes of Ising and Spin-glass models by \cite{becker} and \cite{zhou}, we compute \emph{Energy Landscape Maps} (ELMs) in the high-dimensional model spaces (i.e. the hypothesis spaces in the machine learning literature) for some classic statistical learning problems --- clustering and bi-clustering. \begin{figure} \center \includegraphics[scale=.40]{figures/dg_illustration} \caption{An energy function and the corresponding Energy Landscape Map (ELM). The y-axis of the ELM is the energy level, each leaf node is a local minimum and the leaf nodes are connected at the ridges of their energy basins. } \label{fig:mapping-to-elm} \end{figure} The ELM is a tree structure, as Figure~\ref{fig:mapping-to-elm} illustrates, in which each leaf node represents a local minimum and each non-leaf node represents the barrier between adjacent energy basins. The ELM characterizes the energy landscape with the following information. \begin{itemize} \item The number of local minima and their energy levels; \item The energy barriers between adjacent local minima and their energy levels; and \item The probability mass and volume of each local minimum (See Figure~\ref{fig:2d-mass-vol}). \end{itemize} Such information is useful in the following tasks. \begin{enumerate} \item Analyzing the intrinsic difficulty (or complexity) of the optimization problems, for either inference or learning tasks. For example, in bi-clustering, we divide the problem into the {\em easy}, {\em hard}, and {\em impossible} regimes under different conditions. \item Anlyzing the effects of various conditions on the ELM complexity, for example, the separability in clustering, the number of training examples, the level of supervision (i.e. how many percent the examples are labeled), and the strength of regularization (i.e. prior model). \item Analyzing the behavior of various algorithms by showing their frequencies of visiting the various minima. For example, in the muilti-Gaussian clustering problem, we find that when the Gaussian components are highly separable, K-means clustering works better than the EM algorithm \cite{rubin}, and the opposite is true when the components are less separable. In contrast to the frequent visits of local minimum by K-means and EM, the Swendsen-Wang cut method \cite{zhu-sw-cut} converges to the global minimum in all separability conditions. \end{enumerate} \begin{figure}[htbp!] \center \subfigure[] { \includegraphics[width=0.6\columnwidth]{figures/EnergyLanscape2param.png} } \subfigure[]{ \includegraphics[width=0.2\columnwidth]{figures/ELM2param.png} } \caption{ (a) Energy Landscape for a 4-component 1-d GMM with all parameters fixed except two means. Level sets are highlighted in red. The local minima are shown in yellow dots and the first 200 MCMC samples are shown in black dots. (b) The resulting ELM and the correspondence between the leaves and the local minima from the energy landscape.} \label{fig:2vars} \end{figure} \begin{figure} [htbp!] \center \subfigure[probability mass] { \includegraphics[width=0.25 \columnwidth]{figures/2d-mass.png} } \hspace{2cm} \subfigure[volume]{ \includegraphics[width=0.25 \columnwidth]{figures/2d-volume.png} } \caption{The probability mass and volume of the energy basins for the 2-d landscape shown in Figure \ref{fig:2vars}. } \label{fig:2d-mass-vol} \end{figure} We start with a simple illustrative example in Figures~\ref{fig:2vars} and \ref{fig:2d-mass-vol}. Suppose the underlying probability distribution is a 4-component Gaussian mixture model (GMM) in 1D space, and the components are well separated. The model space is 11-dimensional with parameters $\{ (\mu_i, \sigma_i, \alpha_i): i=1,2,3,4\}$ denoting the means, variance and weights for each components. We sampled $70$ data points from the GMM and construct the ELM in the model space. We bound the model space to a finite range defined by the samples. As we can only visualize 2D maps, we set all parameters to equal the truth value except keeping $\mu_1$ and $\mu_2$ as the unknowns. Figure \ref{fig:2vars}(a) shows the energy map on a range of $0 \leq \mu_1, \mu_2 \leq 5$. The asymmetry in the landscape is caused by the fact that the true model has different weights between the first and second component. Some shallow local minima, like E, F, G,H, are little ``dents'' caused by the finite data samples. Figure \ref{fig:2vars} (a) shows that all the local minima are identified. Additionally, it shows the first 200 MCMC samples that were accepted by the algorithm that we will discuss late. The samples are clustered around the local minima, and cover all energy basins. They are not present in the high energy areas away from the local minima, as would be desired. Figure \ref{fig:2vars} (b) shows the resulting ELM and the correspondence between the leaves and the local minima in the energy landscape. Furthermore, Figures \ref{fig:2d-mass-vol} (a) and (b) show the probability mass and the volume of these energy basins. In the literature, \cite{becker} presents the first work for visualizing multidimensional energy landscapes for the spin-glass model. Since then statisticians have developed a series of MCMC methods for improving the efficiency of the sampling algorithms traversing the state spaces. Most notably, \cite{WL_Liang,GWL} generalize the Wang-Landau algorithm~\cite{WL} for random walks in the state space. \cite{zhou} uses the generalized Wang-Landau algorithm to plot the disconnectivity graph for Ising model with 100s of local minimum and proposes an effective way for estimating the energy barrier. Furthermore, \cite{Zhou2} construct the energy landscape for Bayesian inference of DNA sequence segmentation by clustering Monte Carlo samples. In contrast to the above work that compute the landscapes in ``state'' spaces for inference problems, our work is focused on the landscapes in ``model'' spaces (the sets of all models; also called hypothesis spaces in the machine learning community) for statistical learning and model estimation problems. There are some new issues in plotting the model space landscapes. i) Many of the basins have a flat bottom, for example, basin A in Figure~\ref{fig:2vars}.(a). This may result in a large number of false local minima. ii) The are constraints among some parameters, for example the weights have to sum to one --- $\sum_i \alpha_i =1$. Thus we may need to run our algorithm on a manifold. \section{ELM construction} In this section, we introduce the basic ideas for constructing the ELM and estimating its properties \-- mass, volume and complexity. \subsection{Space partition} Let $\Omega$ be the model space over which a probability distribution $\pi(x)$ and energy $E(x)$ are defined. In this paper, we assume $\Omega$ is bounded using properties of the samples. $\Omega$ is partitioned into $K$ disjoint subspaces which represent the energy basins \begin{equation} \Omega = \cup^{K}_{i=1} D_i, \quad \cap_{i=1}^K D_i = \emptyset ~~\forall i\neq j. \end{equation} That is, any point $x \in D_i$ will converge to the same minimum through gradient descent. \begin{figure} \center \includegraphics[width= 0.4 \columnwidth] {figures/binning.jpg} \caption{ The model space $\Omega$ is partitioned into energy basins $D_i$ (along the x-axis), and the energy $\mathbb{R}$ (the y-axis) is partitioned into uniform intervals $ [u_{j+1}, u_j)$. } \label{fig:binning} \end{figure} As Figure~\ref{fig:binning} shows, the energy is also partitioned into intervals $ [u_{j+1}, u_j), j=1,2,...,L$. Thus we obtain a set of bins as the quantized atomic elements in the product space $\Omega \times \mathbb{R}$, \begin{equation} B_{ij} = \{ x: x \in D_i, ~ E(x) \in [u_{j+1}, u_j)\}. \end{equation} The number of basins $K$ and the number of intervals $L$ are unknown and have to be estimated during the computing process in an adaptive and iterative manner. \subsection{Generalized Wang-Landau algorithm} The objective of the generalized Wang-Landau (GWL) algorithm is to simulate a Markov chain that visits all the bins $\{ B_{ij}, \forall i,j\}$ with equal probability, and thus effectively reveal the structure of the landscape. Let $\phi: \Omega \rightarrow \{1,\dots,K\} \times \{1, ...,L\}$ be the mapping between the model space and bin indices: $\phi(x) =(i,j)$ if $x\in B_{ij}$. Given any $x$, by gradient descent or its variants, we can find and record the basin $D_i$ that it belongs to, compute its energy level $E(x)$, and thus find the index $\phi(x)$. We define $\beta(i,j)$ to be the probability mass of a bin \begin{equation} \beta(i,j) = \int_{B_{i,j}} \pi(x) ~ dx. \end{equation} Then, we can define a new probability distribution which has equal probability among all the bins, \begin{equation} \pi'(x) = \frac{1}{Z} \pi(x) / \beta(\phi(x)), \end{equation} with $Z$ being a scaling constant. To sample from $\pi'(x)$, one can estimate $\beta(i,j)$ by a variable $\gamma_{ij}$. We define the probability function $\pi_\gamma: \Omega \rightarrow \mathbb{R}$ to be \begin{align*} \pi_\gamma(x) \propto \frac{\pi(x) }{\gamma_{\phi(x)}} = \sum_{i,j} \frac{\pi(x)}{\gamma_{ij}} \mathbbm{1}(x \in B_{ij}) ~\text{ st. } \int_{\Omega} \pi_\gamma(x) dx = 1. \end{align*} We start with an initial $\gamma^0$, and update $\gamma^t =\{\gamma_{ij}^t, \forall i,j\}$ iteratively using stochastic approximation \cite{Stoch_grad}. Suppose $x_t$ is the MCMC state at time $t$, then $\gamma^t$ is updated in an exponential rate, \begin{equation} \label{eq:update} \log \gamma_{ij}^{t+1} = \log \gamma_{ij}^{t} + \eta_t \mathbbm{1}(x_t \in B_{ij}), \quad \forall i,j. \end{equation} $\eta_t$ is the step size at time $t$. The step size is decreased over time and the decreasing schedule is either pre-determined as in \cite{Stoch_grad} or determined adaptively as in \cite{zhou2011multi}. Each iteration with given $\gamma^t$ uses a Metropolis step. Let $Q(x,y)$ be the proposal probability for moving from $x$ to $y$, then the acceptance probability is \begin{eqnarray}\label{eq:acceptance} \alpha(x,y) &= \min\left(1, \frac{Q(y,x) \pi_{\gamma}(y)}{Q(x,y) \pi_{\gamma}(x)}\right)\\ \nonumber & = \min\left( 1, \frac{Q(y,x)}{Q(x,y)} \frac{\pi(y)} {\pi(x)} \frac{\gamma^t_{\phi(x)}}{\gamma^t_{\phi(y)}}\right). \end{eqnarray} Intuitively, if $\gamma^t_{\phi(x)} < \gamma^t_{\phi(y)}$, then the probability of visiting $y$ is reduced. For the purpose of exploring the energy landscape, the GWL algorithm improves upon conventional methods, such as the simulated annealing \cite{geyer} and tempering \cite{marinari} process. The latter sample from $\pi(x)^{\frac{1}{T}}$ and do not visit the bins with equal probability even at high temperature. \begin{figure} \center { \includegraphics[width = 0.4 \columnwidth ]{figures/gradient_descent2.pdf} } \caption{First two steps of projected gradient descent. The algorithm is initialized with MCMC sample $x_t$. $v$ is the gradient of $E(x)$ at the point $x_t$. Armijo line search is used to determine the step size $\alpha$ along the vector $v$. $x'_t$ is the projection $T(x_t+\alpha v)$ onto the subspace $\Gamma$. Then $x''_t$ is the projection $T(x_t+\alpha' v')$, and so on.} \label{fig:projection} \end{figure} In performing gradient descent, we employ Armijo line search to determine the step size; if the model space $\Omega$ is a manifold in $\mathbb{R}^n$, we perform projected gradient descent, as shown in Figure \ref{fig:projection}. To avoid erroneously identifying multiple local minima within the same basin (especially when there is large flat regions), we merge local minima identified by gradient descent based on the following criteria: (1) the distance between two local minima is smaller than a constant $\epsilon$; or (2) there is no barrier along the straight line between two local minima. Figure \ref{fig:paths} (a) illustrates a sequence of Markov chain states $x_t, ..., x_{t+9}$ over two energy basins. The dotted curves are the level sets of the energy function. \begin{figure}[t] \center \includegraphics[scale=.45]{figures/MCMC_path3.pdf} \caption{Sequential MCMC samples $x_t, x_{t+1}, \dots, x_{t+9}$. For each sample, we perform gradient descent to determine which energy basin the sample belongs to. If two sequential samples fall into different basins ($x_{t+3}$ and $x_{t+4}$ in this example), we estimate or update the upper-bound of the energy barrier between their respective basins ($B_1$ and $B_2$ in this example).} \label{fig:paths} \end{figure} \subsection{Constructing the ELM} Suppose we have collected a chain of samples $x_1, \dots, x_N$ from the GWL algorithm. The ELM construction consists of the following two processes. \begin{figure} \center \includegraphics[width = 0.5\columnwidth]{figures/ridge_descent.pdf} \caption{The ridge descent algorithm is used for estimating the energy barrier between basins $D_k$ and $D_l$ initialized at consecutive MCMC samples $a_0 = x_t, b_0 = x_{t+1}$ where $a_0 \in D_k$ and $b_0 \in D_l$. } \label{fig:ridge} \end{figure} {\em 1, Finding the energy barriers between adjacent basins}. We collect all consecutive MCMC states that move across two basins $D_k$ and $D_l$, \begin{equation} X_{kl} = \{ (x_t, x_{t+1}): x_t \in D_k, x_{t+1} \in D_l\} \end{equation} we choose $(a_0,b_0) \in X_{kl}$ with the lowest energy \[(a_0,b_0) = \text{argmin}_{(a,b) \in \Omega_{kl}}\left[\min(E(a),E(b))\right]. \] Next we iterate the following step as Figure~\ref{fig:ridge} illustrates \begin{align*} a_i &= \text{argmin}_a \left\{ E(a): a\in\text{Neighborhood}(b_{i-1}) \cap D_k \right\} \\ b_i &= \text{argmin}_b \left\{ E(b): b\in\text{Neighborhood}(a_{i}) \cap D_l \right\} \end{align*} until $b_{i-1} = b_i$. The neighborhood is defined by an adaptive radius. Then $b_i$ is the energy barrier and $E(b_i)$ is the energy level of the barrier. A discrete version of this ridge descent method was used in \cite{zhou}. {\em 2, Constructing the tree structure}. The tree structure of the ELM is constructed from the set of energy basins and the energy barriers between them via an iterative algorithm modified from the hierarchical agglomerative clustering algorithm. Initially, the energy basins are represented by leaf nodes that are not connected, whose y-coordinates are determined by the local minima of the basins. In each iteration, the two nodes representing the energy basins $D_1, D_2$ with the lowest barrier are connected by a new parent node, whose y-coordinates is the energy level of the barrier; $D_1$ and $D_2$ are then regarded as merged, and the energy barrier between the merged basin and any other basin $D_i$ is simply the lower one of the energy barriers between $D_1/D_2$ and $D_i$. When all the energy basins are merged, we obtain the complete tree structure. For clarity, we can remove from the tree basins of depth less than a constant $\epsilon$. \subsection{Estimating the mass and volume of nodes in the ELM}\label{sec:alg} In the ELM, we can estimate the probability mass and the volume of each energy basin. When the algorithm converges, the normalized value of $\gamma_{ij}$ approaches the probability mass of bin $B_{ij}$: \[ \hat{P}(B_{ij}) = \frac{\gamma_{ij}}{\sum_{kl} \gamma_{kl}} \rightarrow \beta(i,j), \quad {\rm almost \; surely}. \] Therefore the probability mass of a basin $D_i$ can be estimated by \begin{equation} \hat{P}(D_i) = \sum_j \hat{P}(B_{ij}) = \frac{\sum_j \gamma_{ij}}{\sum_{kl} \gamma_{kl}} \end{equation} Suppose the energy $E(x)$ is partitioned into sufficiently small intervals of size $du$. Based on the probability mass, we can then estimate the size\footnote{Note that the size of a bin/basin in the model space is called its volume by \cite{Zhou2}, but here we will use the term ``volume'' to denote the capacity of a basin in the energy landscape.} of the bins and basins in the model space $\Omega$. A bin $B_{ij}$ with energy interval $[u_j, u_j+du)$ can be seen as having energy $u_j$ and probability density $\alpha e^{-u_j}$ ($\alpha$ is a normalization factor). The size of bin $B_{ij}$ can be estimated by \[ \hat{A}(B_{ij}) = \frac{\hat{P}(B_{ij})}{\alpha e^{-u_j}} = \frac{\gamma_{ij}}{\alpha e^{-u_j} \sum_{kl} \gamma_{kl}} \] The size of basin $D_i$ can be estimated by \begin{equation} \hat{A}(D_i) = \sum_j \hat{A}(B_{ij}) = \frac{1}{\sum_{kl} \gamma_{kl}} \sum_j \frac{\gamma_{ij}}{\alpha e^{-u_j}} \end{equation} Further, we can estimate the volume of a basin in the energy landscape which is defined as the amount of space contained in a basin in the space of $\Omega\times \mathbb{R}$. \begin{equation} \hat{V}(D_i) = \sum_{j} \sum_{k: u_k \leq u_j} \hat{A}(B_{ik}) \times du = \frac{du}{\sum_{lm} \gamma_{lm}} \sum_{j} \sum_{k: u_k \leq u_j} \frac{\gamma_{ik}}{\alpha e^{-u_k}} \end{equation} where the range of $j$ depends on the definition of the basin. In a restricted definition, the basin only includes the volume under the closest barrier, as Figure~\ref{fig:mass-volume-calc} illustrates. The volume above the basins $1$ and $2$ is shared by the two basins, and is between the two energy barriers $C$ and $D$. Thus we define the volume for a non-leaf node in the ELM to be the sum of its childen plus the volume between the barriers. For example, node $C$ has volume $V(A) + V(B) + V(AB)$. \begin{figure}[t] \center \includegraphics[width= 0.8 \columnwidth] {figures/ELM_volume_illustration.png} \caption{The volume of basins. Assuming that $du$ is sufficiently small, the volume of an energy basin can be approximated by the summation of the estimated volume at each energy interval. } \label{fig:mass-volume-calc} \end{figure} If our goal is to develop a scale-space representation of the ELM by repeatedly smoothing the landscape, then basins $A$ and $B$ will be merges into one basin at certain scale, and volume above the two basins will be also added to this new merged basin. Note that the partition of the space into bins, rather than basins, facilitates the computation of energy barriers, the mass and volume of the basins. \begin{figure}[t] \center \includegraphics[width= 1.0 \columnwidth] {figures/Difficulty_measure.PNG} \caption{Characterizing the difficulty of learning in the ELM. For two learning tasks with ELM I and ELM II, the colored bar show the frequency that a learning algorithm converges to the basins, from which two Error-recall curves are plotted. The difficulty of learning task, with respect to this algorithm, can be measured by the area under the curve within an acceptable maximum error.} \label{fig:difficulty_measure} \end{figure} \subsection{Characterizing the difficulty (or complexity) of learning tasks \label{sect:difficulty}} It is often desirable to measure the difficulty of the learning task by a single number. For example, we compare two ELMs in Figure~\ref{fig:difficulty_measure}. Learning in the landscape of ELM I looks easier than that of ELM II. However, the difficulty also depends on the learning algorithms. Thus we can run the learning algorithm many times and record the frequency that it converges to each basin or minimum. The frequency is shown by the lengths of the colored bars under the leaf nodes. Suppose that $\Theta^\ast$ is the true model to be learned. In Figure~\ref{fig:difficulty_measure}, $\Theta^\ast$ corresponds to nodes $X$ in ELM I and node $A$ in ELM II. In general, $\Theta^\ast$ may not be the global minimum or not even a minimum. We then measure the distance (or error) between $\Theta^\ast$ and any other local minima. As the error increases, we accumulate the frequency to plot a curve. We call it the Error-Recall curve (ERC), as the horizontal axis is the error and the vertical axis is the frequency of recall the solutions. This is like the ROC (receptor-operator characteristics) curves in Bayesian decision theory, pattern recognition and machine learning. By sliding the threshold $\epsilon_{\rm max}$ which is maximum error tolerable, the curve characterizes the difficulty of the ELM with respect to the algorithm. A single numeric number that characterizes the difficulty can be the area under the curve (AUC) for a given $\epsilon_{\rm max}$. this is illustarted by the shadowed area in \ref{fig:difficulty_measure}.(c) for ELM II. When $AUC$ is close to $1$, the task is easy, and when $AUC$ is close to $0$, learning is impossible. In a learning problem, we can set different conditions which correspond to a range of ELMs. The difficulty measures of these ELMs can be visualized in the space of the parameters as a difficulty map. We will show such maps in experiment III. \subsection{MCMC moves in the model space} To design the Markov chain moves in the model space $\mathbb{R}$, we use two types of proposals in the metropolis-Hastings design in equation~(\ref{eq:acceptance}). 1, A random proposal probability $Q(x,y)$ in the neighborhood of the current model $x$. 2, Data augmentation. A significant portion of non-convex optimization problems involve latent variables. For example, in the clustering problem, the class label of each data point is latent. For such problems, we use data augmentation \cite{tanner1987calculation} to improve the efficiency of sampling. In order to propose a new model $y=x_{t+1}$, we first sample the values of the latent variables $Z_t$ based on $p(Z_t|x_t)$ and then sample the new model $x_{t+1}$ based on $p(x_{t+1}|Z_t)$. The proposal $y=x_{t+1}$ is then either accepted or rejected based on the same acceptance probability in Equation~\ref{eq:acceptance}. Note that, however, our goal in ELM construction is to traverse the model space instead of sampling from the probability distribution. When enough samples are collected and therefore the weights $\gamma_{ij}$ become large, the reweighted probability distribution would be significantly different from the original distribution $\pi(x)$ and the rejection rate of the models proposed via data augmentation would become high. Therefore, we use the proposal probability based on data augmentation more often at the beginning and increasingly rely on random proposal when the weights become large. \begin{figure} \center \includegraphics[width=0.6\columnwidth]{figures/two_trees_convergence.png} \caption{ Two ELMs generated from two MCMC chains $C_1$ and $C_2$ initialized at different starting points after convergence in $24,000$ iterations. } \label{fig:ELM_variations} \end{figure} \subsection{ELM convergence analysis} The convergence of the GWL algorithm to a stationary distribution is a necessary but not sufficient condition for the convergence of the ELMs. As shown in Figure~\ref{fig:ELM_variations}, the constructed ELMs may have minor variations due to two factors: (i) the left-right ambiguity when we plot the branches under a barrier; and (ii) the precision of the energy barriers will affect the internal structure of the tree. In experiments, firstly we monitor the convergence of the GWL in the model space. We run multiple MCMC initialized with random starting values. After a burn-in period, we collect samples and project in a 2-3 dimensional space using Multi-dimensional scaling. We check whether the chains have converged to a stationary distribution using the multivariate extension of the Gelman and Rubin criterion~\cite{Gelman1}\cite{Gelman2}. Once the GWL is believed to have converged, we can monitor the convergence of the ELM by checking the convergence of the following two sets over time $t$. \begin{enumerate} \item The set of leaf notes of the tree $S_L^t$ in which each point $x$ is a local minimum with energy $E(x)$. As $t$ increase, $S_L^t$ grows monotonically until no more local minimum is found, as is shown in Figure~\ref{fig:convergence}.(a). \item The set of internal nodes of the tree $S_N^t$ in which each point $y$ is an energy barrier at level $E(y)$. As $t$ increases, we may find lower barrier as the Markov chain crosses different ridge between the basins. Thus $E(y)$ decreases monotonically until no barrier in $S_N^t$ is updated during a certain time period. \end{enumerate} We further calculate a distance measure between two ELMs constructed by two MCMCs with different initialization. To do so, we compute a best node matching between the two trees and then the distance is defined on the differences of the matched leaf nodes and barriers, and penalties on unmatched nodes. We omit the details of this definition as it is not important for this work. Figure~\ref{fig:convergence}.(b) shows the distance decreases as more samples are generated. \begin{figure} \center \subfigure[] { \includegraphics[width=0.38\columnwidth]{figures/chainConvergence2.png} } \subfigure[] { \includegraphics[width=0.5\columnwidth]{figures/error_decrease.png} } \caption{ Monitoring the convergence of ELMs generated from two MCMC chains $C_1$ and $C_2$ initialized at different starting points. (a) The number of local minima found vs number of iterations for $C_1$ and $C_2$. (b) the distance between the two ELMs vs. number of iterations.} \label{fig:convergence} \end{figure} \section{Experiment I: ELMs of Gaussian Mixture Models}\label{sec:gmm} In this section, we compute the ELMs for learning Gaussian mixture models for two purpuses: (i) study the influences of different conditions, such as separability and level of supervision; and ii) compare the behaviors and performances of popular algorithms including K-mean clustering, EM (Expectation-Maximization), two-step EM, and Swendson-Wang cut. We will use both synthetic data and real data in the experiments. \subsection{Energy and Gradient Computations} A Gaussian mixture model $\Theta$ with $n$ components in $d$ dimensions have weights $\alpha_i$, means $\mu_i$ and covariance matrices $\Sigma_i$ for $i=1,\dots,n$. Given a set of observed data points $\{z_i, i=1,...,m\}$, we write the energy function as \begin{eqnarray}\label{eq:gmm_energy} E(\Theta) &=& - \log P( z_i : i=1\dots m | \Theta ) - \log P(\Theta) \\ &=& - \sum^m_{i=1} \log f(z_i | \Theta) - \log P(\Theta). \end{eqnarray} $P(\Theta)$ is the product of a Dirichlet prior and a NIW prior. Its partial derivatives are trivial to compute. $f(z_i | \Theta) = \sum^n_{j=1} \alpha_j G(z_i; \mu_j, \Sigma_j)$ is the likelihood for data $z_i$, where $G(z_i; \mu_j, \Sigma_j) = \frac{1}{{ \sqrt {\det(2\pi \Sigma_j )} }} \exp \left[-\frac{1}{2}\left( {z_i - \mu_j } \right)^T \Sigma_j^{-1} \left( {z_i - \mu_j } \right)\right]$ is a Gaussian model. For a sample $z_i$, we have the following partial derivatives of the log likelihood for calculating the gradient in the energy landscape. a), Partial derivative with respect to each weight $\alpha_j$: \begin{align*} \frac{\delta \log f(z_i) }{\delta \alpha_j} = \frac{ G(z_i; \mu_j, \Sigma_j)} {\sum^K_{k=1} \alpha_k G(z_i, \mu_k, \Sigma_k)}. \end{align*} b), Partial derivative with respect to each mean $\mu_j$: \begin{align*} \frac{\delta \log f(z_i) }{\delta \mu_j} = \frac{\alpha_j G(z_i; \mu_j, \Sigma_j)}{ \sum^K_{k=1} \alpha_k G(z_i; \mu_k, \Sigma_k)} \Sigma_j^{-1}(\mu_j - z_i ). \end{align*} c), Partial derivative with respect to each covariance $\Sigma_j$: \begin{align*} \frac{\delta \log f_\text{mm}(z_i) }{\delta \Sigma_j} & = \frac{\alpha_j G(z_i; \mu_j, \Sigma_j)} { \sum^K_{k=1} \alpha_k G(z_i; \mu_k,\Sigma_k)} \frac{1}{2} \left[ \frac{\delta}{\delta \Sigma_j}\log \alpha_j G(z_i; \mu_j, \Sigma_j) \right] \\ & = \frac{\alpha_j G(z_i; \mu_j, \Sigma_j)} { \sum^K_{k=1} \alpha_k G(z_i; \mu_k,\Sigma_k)} \frac{1}{2} \left[ -\Sigma_j^{-T} + \Sigma_j^{-T} \left( {z_i - \mu_j } \right) \left( {z_i - \mu_j } \right)^T \Sigma_j^{-T} \right] \end{align*} During the computation, we need to restrict the $\Sigma_j$ matrices so that each inverse $\Sigma_j^{-1}$ exists in order to have a defined gradient. Each $\Sigma_j$ is semi-positive definite, so each eigenvalue is greater than or equal to zero. Consequently we only need the minor restriction that for each eigenvalue $\lambda_i$ of $\Sigma_j$, $\lambda_i > \epsilon$ for some $\epsilon > 0$. However, it is possible that after one gradient descent step, the new GMM parameters will be outside of the valid GMM space, i.e. the new $\Sigma^{t+1}_j$ matrices at step $t+1$ will not be symmetric positive definite. Therefore, we need to project each $\Sigma^{t+1}_j$ into the symmetric positive definite space with the projection \begin{align*} P_{\text{symm}}(P_{\text{pos}}(\Sigma^{t+1}_j)). \end{align*} The function $P_{\text{symm}}(\Sigma)$ projects the matrix into the space of symmetric matrices by \begin{align*} P_{\text{symm}}(\Sigma) = \frac{1}{2} (\Sigma+ \left(\Sigma\right)^T). \end{align*} Assuming that $\Sigma$ is symmetric, the function $P_{\text{pos}}(\Sigma)$ projects $\Sigma$ into the space of symmetric matrices with eigenvalues greater than $\epsilon$. Because $\Sigma$ is symmetric, it can be decomposed into $\Sigma = Q \Lambda Q^T$ where $\Lambda$ is the diagonal eigenvalue matrix $\Lambda = diag\{\lambda_{1}, \dots, \lambda_{n}\}$, and $Q$ is an orthonormal eigenvector matrix. Then the function \begin{align*} P_{\text{pos}}(\Sigma) = Q \begin{pmatrix} \max(\lambda_{1},\epsilon) & 0 & \ldots & 0\\ 0 & \max(\lambda_{2},\epsilon) & \ldots & 0\\ \vdots & \vdots & \ddots & \vdots\\ 0 & 0 &\ldots & \max(\lambda_{n},\epsilon) \end{pmatrix} Q^T \end{align*} ensures that $P_{\text{pos}}(\Sigma)$ is symmetric positive definite. \subsection{Bounding the GMM space} From the $m$ data points $\{z_i, i = 1,\dots, m\}$, we can estimate a boundary of the space of possible parameter $\Theta$ if $m$ is sufficiently large. Let $\mu_o$ and $\Sigma_o$ be the sample mean and sample covariance matrix of all $m$ points. We set a range for the means $\mu_j$ of the Gaussian components, \begin{align*} ||\mu_j - \mu_o ||_2 < \max_i || z_i - \mu_o||_2 + \epsilon_m. \end{align*} $\epsilon_m$ is a constant that we will select in experiments. To bound the covariance matrices $\Sigma_j$, let $\Sigma_o = Q \Lambda Q^T$ be the eigenvalue decomposition of $\Sigma_o$ with $\Lambda =diag\{\lambda_1, \cdots, \lambda_n\}$. We denote by $L = \max(\lambda_1, \dots, \lambda_n) + \epsilon_m$ the upper bound of the eigen-values, and bound all the eigenvalues of $\Sigma_j$ by $L$. \begin{figure} \center \subfigure[unbounded GMM space] { \includegraphics[width=0.45\columnwidth]{figures/cluster_centers_unbounded.png} } \subfigure[bounded GMM space]{ \includegraphics[width=0.38\columnwidth]{figures/cluster_centers.png} } \caption{We sampled 70 data points from a 1-dimensional 4-component GMM and ran the MCMC random walk for ELM construction algorithm in the (a) unbounded (b) bounded GMM space. The plots show the evolution of the location of the centers of the 4 components over time. The width of the line represents the weight of the corresponding component. } \label{fig:cluster-centers} \end{figure} Figure \ref{fig:cluster-centers} (a,b) compare the MCMCs in unbounded and bounded spaces repsectively. We sampled $m=70$ data points from a 1-dimensional, 4-component GMM and ran the MCMC random walk for ELM construction algorithm. The plots show the evolution of the locations of $\mu_1,...,\mu_4$ over time. Notice that in Figure \ref{fig:cluster-centers} (a), the MCMC chain can move far from the center and spends the majority of the time outside of the bounded subspace. In Figure \ref{fig:cluster-centers} (b), by forcing the chain to stay within the boundary, we are able to explore the relevant subspace more efficiently. \subsection{Experiments on Synthetic Data} We start with synthetic data with $k=3$ component GMM on $2$ dimensional space, draw $m$ samples and run our algorithm to plot the ELM under different settings. \begin{figure} \center \includegraphics[width= \columnwidth]{figures/trees-by-sep-mass.png} \caption{ELMs for $m=100$ samples drawn from GMMs with low, medium and high separability $c = 0.5, 1.5, 3.5$ respectively. The circle represents the probability mass of the basins. } \label{fig:trees-by-sep} \end{figure} {\bf 1) The effects of separability}. The separability of the GMM represents the overlap between components of the model and is defined as $c = \min \left( \frac{||\mu_i - \mu_j|| }{\sqrt{n}\max(\sigma_1,\sigma_2)} \right)$. This is often used in the literature to measure the difficulty of learning the true GMM model. Figure~\ref{fig:trees-by-sep} shows three representative ELMs with the separability $c=0.5, 1.5, 3.5$ respectively for $m = 100$ data points. This clearly shows that at $c=0.5$, the model is hardly identifiable with many local minima reaching similar energy levels. The energy landscape becomes increasingly simple as the separability increases. When $c = 3.5$, the prominent global minimum dominates the landscape. {\bf 2) The effects of partial supervision}. We assign ground truth labels to a portion of the $m$ data points. For $z_i$, its label $\ell_i$ indicates which component it belongs to. We set $m=100$, separability $c=1.0$. Figure \ref{fig:partial-lab-trees} shows the ELMs with $0\%, 5\%, 10\%, 50\%, 90\%$ data points labels. While unsupervised learning ($0\%$) is very challenging, it becomes much simpler when $5\%$ or $10\%$ data are labeled. When $90\%$ data are labeled, the ELM has only one minimum. Figure \ref{fig:lm_by_labels} shows the number of local minima in the ELM when labeling $1, \dots, 100$ samples. This shows a significant decrease in landscape complexity for the first 10\$ labels, and diminishing returns from supervised input after the initial 10\%. \begin{figure} \center \includegraphics[width=\textwidth] {figures/partial_label_trees.png} \caption{ ELMs with of synthesized GMMs (separability $c=1.0$, nSamples = 100) with $\{0\%, 5\%, 10\%, 50\%, 90\% \}$ labelled data points.} \label{fig:partial-lab-trees} \end{figure} \begin{figure} \center \includegraphics[width= 0.5\columnwidth] {figures/lm_by_labels.png} \caption{ Number of local minima versus the percentage of labelled data points for a GMM with separability $c=1.0$. } \label{fig:lm_by_labels} \end{figure} {\bf 3) Behavior of Learning Algorithms.} We compare the behaviors of the following algorithms under different separability conditions. \begin{itemize} \item Expectation-maximization (EM) is the most popular algorithms for learning GMM in statistics. \item K-means clustering is a popular algorithm in machine learning and pattern recognition. \item Two-step EM is a variant of EM proposed in \cite{dasgupta} who have proved a performance guarantee under certain separability conditions. It starts with an excessive number of components and then prune them. \item The Swedsen-Wang Cut (SW-cut) algorithm proposed in \cite{zhu-sw-cut} and \cite{zhu-sw-cut-pami}. This generalizes the SW method \cite{sw-cut} from Ising/Potts models to arbitrary probabilities. \end{itemize} We modified EM, two-step EM and SW-cut in our experiments so that they minimize the energy function defined in Equation \ref{eq:gmm_energy}. K-means does not optimize our energy function, but it is frequently used as an approximate algorithm for learning GMM and therefore we include it in our comparison. For each synthetic dataset in the experiment, we first construct the ELM, and then ran each of the algorithms for $200$ times and record which of the energy basins the algorithm lands to. Hence we obtain the visiting frequency of the basins by each algorithm, which are shown as bars of varying length at the leaf nodes in Figures~\ref{fig:scatteredResults} and \ref{fig:behavior}. \begin{figure} \center \includegraphics[width=\columnwidth]{figures/scatteredResults.png} \caption{ The performance of the k-means, EM and 2-step EM algorithms on the ELMs with 10 samples drawn from a GMM with low separability ($c = 0.5$) } \label{fig:scatteredResults} \end{figure} \begin{figure} \center \subfigure[EM] { \includegraphics[width=0.30\textwidth,height=150pt]{figures/trees/EM-low-sep.png} } \subfigure[k-means]{ \includegraphics[width=0.30\textwidth,height=150pt]{figures/trees/kmeans-low-sep.png} } \subfigure[SW-cut]{ \includegraphics[width=0.33\textwidth,height=150pt]{figures/trees/SW-low-sep.png} } \subfigure[EM] { \includegraphics[width=0.27\textwidth,height=120pt]{figures/trees/EM-high-sep.png} } \subfigure[k-means]{ \includegraphics[width=0.33\textwidth,height=120pt]{figures/trees/kmeans-high-sep.png} } \subfigure[SW-cut]{ \includegraphics[width=0.33\textwidth,height=120pt]{figures/trees/SW-high-sep.png} } \caption{ The performance of the EM, k-means, and SW-cut algorithm on the ELM. (a-c) Low separability $c = 0.5$. (d-f) High separability $c = 3.5$. } \label{fig:behavior} \end{figure} Figure~\ref{fig:scatteredResults} shows a comparison between the K-means, EM and two-step EM algorithms for $n=10$ samples drawn from a low ($c=0.5$) separability GMM. The results are scattered across different local minima regardless of the algorithm. This illustrates the difficulty in learning a model from a landscape with many local minima separated by large energy barriers. Figure~\ref{fig:behavior} show a comparison of the EM, k-means, and SW-cut algorithms for $m=100$ samples drawn from low ($c = 0.5$) and high ($c = 3.5$) separability GMMs. The SW-cut algorithm performs best in each situation, always converging to the global optimal solution. In the low separability case, the k-means algorithm is quite random, while the EM algorithm almost always finds the global minimum and thus outperforms k-means. However, in the high separability case, the k-means algorithm converges to the true model the majority of the time, while the EM almost always converges to a local minimum with higher energy than the true model. This result confirms a recent theoretical result showing that the objective function of hard-EM (with k-means as a special case) contains an inductive bias in favor of high-separability models \cite{Tu12,samdani2012unified}. Specifically, we can show that the actual energy function of hard-EM is: \[ E(\Theta) = -\log P(\Theta|Z) + \min_q \left( \textbf{KL}(q(L)||P(L|Z,\Theta)) + H_q(L) \right) \] where $\Theta$ is the model parameters, $Z={z_1,\ldots,z_m}$ is the set of observable data points, $L$ is the set of latent variables (the data point labels in a GMM), $q$ is an auxiliary distribution of $L$, and $H_q$ is the entropy of $L$ measured with $q(L)$. The first term in the above formula is the standard energy function of clustering with GMM. The second term is called a posterior regularization term \cite{Ganchev10}, which essentially encourages the distribution $P(L|Z,\Theta)$ to have a low entropy. In the case of GMM, it is easy to see that a low entropy in $P(L|Z,\Theta)$ implies high separability between Gaussian components. \subsection{Experiments on Real Data} We ran our algorithm to plot the ELM for the well-known Iris data set from the UCI repository \cite{iris}. The Iris data set contains $150$ points in 4 dimensions and can be modeled by a 3-components GMM. The three components each represent a type of iris plant and the true component labels are known. The points corresponding to the first component are linearly separable from the others, but the points corresponding to the remaining two components are not linearly separable. Figure \ref{fig:iris} shows the ELM of the Iris dataset. We visualize the local minima by plotting the ellipsoids of the covariance matrices centered at the means of each component in 2 of the 4 dimensions. The 6 lowest energy local minima are shown on the right and the 6 highest energy local minima are shown on the left. The high energy local minima are less accurate models than the low energy local minima. The local minima (E) (B) and (D) have the first component split into two and the remaining two (non-separable) components merged into one. The local minima (A) and (F) have significant overlap between the 2nd and 3rd components and (C) has the components overlapping completely. The low-energy local minima (G-L) all have the same 1st components and slightly different positions of the 2nd and 3rd components. \begin{figure} \center \includegraphics[width=\textwidth]{figures/irisPlots.png} \caption{ ELM and some of the local minima of the Iris dataset. } \label{fig:iris} \end{figure} We ran the algorithm with $0, 5, 10, 50, 90, 100$ percent of the points with the ground truth labels assigned. Figure \ref{fig:partial-lab-iris} shows the global minimum of the energy landscape for these cases. \begin{figure} \center \includegraphics[width=\textwidth] {figures/iris-all-labels-scatter3.png} \caption{ Global minima for learning from the Iris dataset with 0, 5, 10, 50, 90, and 100\% of the data labeled with the ground truth values. Unlabeled points are drawn in grey and labelled points are colorized in red, green or blue.} \label{fig:partial-lab-iris} \end{figure} \section{Experiment II: ELM of Bernoulli Templates} The synthetic data and Iris data in experiment I are in low dimensional spaces. In this section, we experiment with very high dimensional data for a learning task in computer vision and pattern recognition. The objective is to learn a number of templates ${\rm BT}_k, k=1,...,K$ for object recognition. Figure~\ref{fig:full_faces} illustrates $10$ templates of animal faces. Each template consists of a number of sketches or edges in the image lattice, and is denoted by a Boolean vector ${\rm BT}_k =(s_{k1}, s_{k2}, \dots, s_{kn})$ with $n$ being the number of quantized positions and orientations of the lattice which is typically a large number $100\sim 1000$. $s_{kj}=1$ if there is a sketch at location $j$, and $s_{kj}=0$ otherwise. Images are generated from one of the $K$ templates with noise. Suppose $z_i =(r_{i1}, r_{12}, \dots, r_{in})$ is an image generated from template ${\rm BT}_k$, then $ r_{ij} = s_{kj} $ with probability $p$ and $ r_{ij} = 1- s_{kj}$ with probability $1-p$. Thus we call ${\rm BT}_k, k=1,2...,K$ the Bernoulli templates. For simplicity we assume $p$ is fixed for all the templates and all the locations. The energy function that we use is the negative log of the posterior, given by $E(\Theta) = -\log P(\Theta |z_i : i=1\dots m)$ for $m$ examples $\{ z_i \}_{i=1}^m$. The model parameter $\Theta$ consists of the Boolean vectors ${\rm BT}_k =(s_{k1}, s_{k2}, \dots, s_{kn})$ and the mixture weights $\alpha_k$ for $k=1,...,K$. By assuming a uniform prior we have \begin{equation*} P(\Theta | z_i: 1=1,...,m) =\prod_{i=1}^m \sum_{k=1}^K \alpha_k p^{\sum_{j=1}^n 1(r_{ij} = s_{kj})}(1-p)^{\sum_{j=1}^n 1(r_{ij} \neq s_{kj})}, \end{equation*} In the following we present experiments on synthetic and real data. \begin{figure*} \center \hspace{-5mm}\subfigure[cat] { \includegraphics[width=0.2\textwidth]{figures/face_pics/cat1.png} } \hspace{-5mm} \subfigure[chilchilla]{ \includegraphics[width=0.2\textwidth]{figures/face_pics/chinchilla1.png} } \hspace{-5mm}\subfigure[dog]{ \includegraphics[width=0.2\textwidth]{figures/face_pics/dog1.png} } \hspace{-5mm} \subfigure[elephant]{ \includegraphics[width=0.2\textwidth]{figures/face_pics/elephant1.png} } \hspace{-5mm} \subfigure[goat]{ \includegraphics[width=0.2\textwidth]{figures/face_pics/goat1.png} }\\ \hspace{-5mm} \subfigure[lion]{ \includegraphics[width=0.2\textwidth]{figures/face_pics/lion2.png} } \hspace{-5mm} \subfigure[monkey]{ \includegraphics[width=0.2\textwidth]{figures/face_pics/monkey3.png} } \hspace{-5mm} \subfigure[mouse]{ \includegraphics[width=0.2\textwidth]{figures/face_pics/mouse2.png} } \hspace{-5mm} \subfigure[owl]{ \includegraphics[width=0.2\textwidth]{figures/face_pics/owl2.png} } \hspace{-5mm} \subfigure[rabbit]{ \includegraphics[width=0.2\textwidth]{figures/face_pics/rabbit1.png} } \caption{Bernoulli templates for animal faces. These templates have low overlap and are well separable.} \label{fig:full_faces} \end{figure*} \subsection{Experiments on Synthetic Data} \cite{Barbu_2run_EM} proposes a Two-Round EM algorithm for learning Bernoulli templates with a performance bound that is dependent on the number of components $K$, the Beronouilli template dimension $n$, and noise level $p$. In this experiment, we examine how the ELM of the model space changes with these factors. We discretize the model space by allowing the weights to take values $\alpha_i \in \{0, 0.1, \dots, 1.0\}$. In order to adapt the GWL algorithm to the discrete space, we use coordinate descent in lieu of gradient descent. \begin{figure} \center \subfigure[Sketch dictionary]{ \includegraphics[width=0.22\columnwidth]{figures/face_pics/dictionary.png} } \hspace{10mm} \subfigure[Noisy dog]{ \includegraphics[width=0.3\columnwidth]{figures/dog4.png} } \caption{(a) Quantized dictionary with 18 sketches for each cell in the image lattice. (b) Sample from the dog animal face template with noise level $p=0.1$} \label{fig:sketch_dict} \end{figure} We construct $10$ Bernouilli templates which represent animal faces in Figure \ref{fig:full_faces}. Each animal face is aligned to a grid of $9\times 9$ cells. Each cell may contain up to $3$ sketches. Within a cell, the sketches are quantized to $18$ discrete location and orientations. More specifically, each sketch is a straight line connecting the endpoints or midpoints of the edges of a square cell, and the $18$ possible sketches in a cell are shown in Figure~\ref{fig:sketch_dict}.(a). They can well approximate the detected edges or Gabor sketches from real images. The Bernouilli template can therefore be represented as a $n=9\times 9 \times 18$ dimensional binary vector. Figure~\ref{fig:sketch_dict}.(b) shows a noisy dog face generated with noisy level $p=0.1$. We compute the ELMs of the Bernouilli mixture model for varying numbers of samples $m = 100, 300, \dots, 7000$ and varying noise level $p=0, 0.05, \dots, 0.5, 0.55$. The number of local minima in each energy landscape is tabulated in Figure \ref{fig:full_faces_landscape} (b) and drawn as a heat map in Figure \ref{fig:full_faces_landscape} (a). As expected, the number of local minima increases as the noise level $p$ increases, and decreases as the number of samples decreases. In particular, with no noise, the landscape is convex and with noise $p>0.45$, there are too many local minima and the algorithm does not converge. \begin{figure} \center \subfigure[map]{ \includegraphics[width=0.25 \columnwidth]{figures/face_pics/landscapePvsSamples.png} } \subfigure[number of local minima]{ \includegraphics[width=0.7 \columnwidth]{figures/face_pics/landscapePvsSamplesByTheNumbers.png} } \caption{ ELM complexity for varying values of $p$ and number of samples $m$ in learning the $10$ Bernouilli Templates. } \label{fig:full_faces_landscape} \end{figure} \begin{figure*} \center \hspace{-4mm} \subfigure[]{ \includegraphics[width=0.14\textwidth]{figures/face_pics/high_overlap/mouse1.png} } \hspace{-4mm} \subfigure[]{ \includegraphics[width=0.14\textwidth]{figures/face_pics/high_overlap/mouse2.png} } \hspace{-4mm} \subfigure[]{ \includegraphics[width=0.14\textwidth]{figures/face_pics/high_overlap/mouse3.png} } \hspace{-4mm} \subfigure[]{ \includegraphics[width=0.14\textwidth]{figures/face_pics/high_overlap/mouse4.png} } \hspace{-4mm} \subfigure[]{ \includegraphics[width=0.14\textwidth]{figures/face_pics/high_overlap/mouse5.png} } \hspace{-4mm} \subfigure[]{ \includegraphics[width=0.14\textwidth]{figures/face_pics/high_overlap/mouse6.png} } \hspace{-4mm} \subfigure[]{ \includegraphics[width=0.14\textwidth]{figures/face_pics/high_overlap/mouse7.png} } \\ \hspace{-4mm} \subfigure[]{ \includegraphics[width=0.14\textwidth]{figures/face_pics/high_overlap/mouse8.png} } \hspace{-4mm} \subfigure[]{ \includegraphics[width=0.14\textwidth]{figures/face_pics/high_overlap/mouse9.png} } \hspace{-4mm} \subfigure[]{ \includegraphics[width=0.14\textwidth]{figures/face_pics/high_overlap/mouse10.png} } \hspace{-4mm} \subfigure[]{ \includegraphics[width=0.14\textwidth]{figures/face_pics/high_overlap/mouse11.png} } \hspace{-4mm} \subfigure[]{ \includegraphics[width=0.14\textwidth]{figures/face_pics/high_overlap/mouse12.png} } \hspace{-4mm} \subfigure[]{ \includegraphics[width=0.14\textwidth]{figures/face_pics/high_overlap/mouse13.png} } \hspace{-4mm} \subfigure[]{ \includegraphics[width=0.14\textwidth]{figures/face_pics/high_overlap/mouse14.png} } \\ \hspace{-4mm} \subfigure[]{ \includegraphics[width=0.14\textwidth]{figures/face_pics/high_overlap/mouse15.png} } \hspace{-4mm} \subfigure[]{ \includegraphics[width=0.14\textwidth]{figures/face_pics/high_overlap/mouse16.png} } \hspace{-4mm} \subfigure[]{ \includegraphics[width=0.14\textwidth]{figures/face_pics/high_overlap/mouse17.png} } \hspace{-4mm} \subfigure[]{ \includegraphics[width=0.14\textwidth]{figures/face_pics/high_overlap/mouse18.png} } \hspace{-4mm} \subfigure[]{ \includegraphics[width=0.14\textwidth]{figures/face_pics/high_overlap/mouse19.png} } \hspace{-4mm} \subfigure[]{ \includegraphics[width=0.14\textwidth]{figures/face_pics/high_overlap/mouse20.png} } \caption{Bernoulli templates for mouse faces with high overlap and low separability.} \label{fig:full_faces_overlap} \end{figure*} We repeat the same experiment using variants of a mouse face. We swap out each component of the mouse face (the eyes, ears, whiskers, nose, mouth, head top and head sides) with three different variants. We thereby generate $20$ Bernouilli templates in Figure~\ref{fig:full_faces_overlap}, which have relatively high degrees of overlap. We generated the ELMs of various Bernouilli mixture models containing three of the $20$ templates and noise level $p=0$. In each Bernouilli mixture model, the three templates have different degrees of overlap. Hence we plot the number of local minima in the ELMs versus the degree of overlap as show in Figure \ref{fig:full_faces_convergence}. As expected, the number of local minima increases with the degree of overlap, and there are too many local minima for the algorithm to converge past overlap $c=0.5$. \begin{figure} \center \includegraphics[width=0.7 \columnwidth]{figures/face_pics/convergence_by_overlap.png} \caption{Number of local minima found for varying degrees of overlap in the Bernoulli templates. Each marker corresponds to a Bernoulli mixture model that consists of three of the 20 Bernoullie templates.} \label{fig:full_faces_convergence} \end{figure} \subsection{Experiments on Real Data} We perform the Bernouilli templates experiment on a set of real images of animal faces. We binarize the images by extracting the prominent sketches on a 9x9 grid. Eight Gabor filters with eight different orientations centered in the centers and corners of each cell are applied to the image. The filters with a strong response above a fixed threshold correspond to edges detected in the image; these are mapped to the dictionary of $18$ elements. Thus each animal face is represented as a $18\times9\times9$ dimensional binary vector. The Gabor filter responses on animal face pictures are shown in Figure \ref{fig:gabor}. The binarized animal faces are shown in Figure \ref{fig:deer-sketches}. \begin{figure}[t] \begin{minipage}[b]{0.45\linewidth} \center \includegraphics[width=0.7\columnwidth]{figures/animal_gabor.png} \caption{Animal face images and corresponding binary sketches indicates the existence of a Gabor filter response above a fixed threshold. } \label{fig:gabor} \end{minipage} \hspace{2em} \begin{minipage}[b]{0.45\linewidth} \center \includegraphics[trim=0cm 9cm 17cm 0cm, clip=true, width=1 \columnwidth]{figures/deer-sketches.png} \caption{Deer face sketches binarized from real images.} \label{fig:deer-sketches} \end{minipage} \end{figure} We chose 3 different animal types -- deer, cat and mouse, with an equal number of images chosen from each category (Figure \ref{fig:all-faces}). The binarized versions of these images can be modeled as a mixture of 3 Bernouilli templates - each template corresponding to one animal face type. \begin{figure} \center \includegraphics[width=0.9 \columnwidth]{figures/all-faces.png} \caption{Animal face images of three categories.} \label{fig:all-faces} \end{figure} The ELM is shown in Figure \ref{fig:all-faces-ELM} along with the Bernouilli templates corresponding to three local minima separated by large energy barriers. We make two observations: 1. The templates corresponding to each animal type are clearly identifiable, and therefore the algorithm has converged on reasonable local minima. 2. The animal faces have differing orientations across the local minima (the deer face on in the left-most local minimum is rotated and tilted to the right and the dog face in the same local minimum is rotated and lilted to the left), which explains the energy barriers between them. \begin{figure} \center \includegraphics[width= 0.5\columnwidth]{figures/tree-real-data-full.png} \caption{ELM of the three animal faces dataset (dog, cat, and deer). We show the Bernouilli templates corresponding to three local minima with large energy barriers. } \label{fig:all-faces-ELM} \end{figure} \begin{figure} \center \subfigure[SW-cut] { \includegraphics[width=0.3\columnwidth]{figures/tree-real-data.png} } \subfigure[EM ]{ \includegraphics[width=0.3\columnwidth]{figures/tree-real-data1.png} } \subfigure[k-means ]{ \includegraphics[width=0.3\columnwidth]{figures/tree-real-data3.png} } \caption{ Comparison of SW-cut, k-means, and EM algorithm performance on the ELM of animal face Bernouilli mixture model.} \label{fig:algs-comp-animals} \end{figure} Figure \ref{fig:algs-comp-animals} shows a comparison of the SW-cut, k-means, and EM algorithm performance as a histogram on the ELM of animal face Bernouilli mixture model. The histogram is obtained by running each algorithm 200 times with a random initialization, then finding the closest local minimum in the ELM to the output of the algorithm. The counts of the closest local minima are then displayed as a bar plot next to each local minimum. It can be seen that SW-cut always finds the global minimum, while k-means performs the worst probably because of the high degree of overlap between the sketches of the three types of animal faces. \section{Experiment III: ELM of bi-clustering}\label{sec:bc} bi-clustering is a learning process (see a survey by \cite{Madeira04survey}) which has been widely used in bioinformatics, e.g., finding genes with similar expression patterns under subsets of conditions (\cite{Cheng00, Getz00, Cho04}). It is also used in data mining, e.g., finding people who enjoy similar movies (\cite{Yang02}), and in learning language models by finding co-occurring words and phrases in grammar rules (\cite{Tu08}). Figure~\ref{fig:bi-cluster}.(a) shows a bi-clustering model (with multiplicative coherence) in the form of a three layer And-Or graph. The underlying pattern $S$ has two conjunction factors $a$ and $b$. $a$ can choose from a number of alternative elements $A_1, A_2, O_1, O_2$ at probability $p_1, ..., p_4$ respectively. Similarly, $b$ can choose from elements $O_1, O_2, B_1, B_2$ with probability $q_1, ..., q_4$ respectively. It can be seen that $a$ and $b$ have shared elements $O_1, O_2$. For comparison, we note that the clustering models in experiments I and II can be seen as three-layer Or-And graphs with a mixture (Or-node) on the top and each component is a conjunction of multiple variables. \begin{figure*} \center \includegraphics[width=\textwidth]{figures/bicluster_illustration.png} \caption{ (a) A bi-clustering model. (b) The co-occurrence matrix with the theoretical frequencies of its elements. } \label{fig:bi-cluster} \end{figure*} From data sampled from this model, one can compute a co-occurrence matrix for the two elements chosen by $a$ and $b$, and the theoretical co-occurring frequency is shown in Figure~\ref{fig:bi-cluster}.(b). When only a small number of observations are available, this matrix may have significant fluctuations. There may also be unwanted background elements in the matrix. The goal of bi-clustering is to identify the bi-cluster (one of the two submatrixes in Figure~\ref{fig:bi-cluster}.(b)) from the noisy matrix. Note that this is a simple special case of the bi-clustering problem and in general the matrix may contain many bi-clusters that are not necessarily symmetrical. We denote the bi-cluster to be identified by $\Theta = \langle A,B \rangle$ where $A$ is the set of rows and $B$ is the set of columns of the bi-cluster. Note that the goal of bi-clustering is not to explain all the data but to identify a subset of the data that exhibit certain properties (e.g., coherence). Therefore, instead of using likelihood or posterior probability to define the energy function, we use the following energy function adapted from \cite{Tu08}. \begin{align*} \label{eqn:bi-clusterLG} E(\Theta) = &\left( s\log s + \sum_{x\in A, y\in B}a_{x,y} \log a_{x,y} - \sum_{x\in A} r_x \log r_x - \sum_{y\in B} c_y\log c_y\right) \\ & - \alpha \left(2\sum_{x\in A, y\in B}a_{x,y} - |A| - |B|\right). \end{align*} In the above formula, $a_{x,y}$ is the element at row $x$ and column $y$, $r_x$ is the sum of row $x$, $c_y$ is the sum of column $y$, and $s$ is the total sum of the bi-cluster. The first term in the energy function measures the coherence of the bi-cluster, which reaches its minimal value of 0 if the bi-cluster is perfectly multiplicatively coherent (i.e., the elements are perfectly proportional). The second term corresponds to the prior, which favors bi-clusters that cover more data; the $-|A|-|B|$ term is added to exclude rows and columns that are entirely zero from the bi-cluster. We experimented with synthetic bi-clustering models in which $a$ and $b$ each have $10$ alternative elements. We varied the following factors to generate a set of different models: (i) the levels of overlaps between $a$ and $b$: $0, 1, ..., 10$; and (ii) random background noises at level $p$. We generated $1000$ data points from each model and constructed the matrix. For each data matrix, we ran our algorithm to plot the ELMs with different values of $\alpha$, the strength of the prior. Figure~\ref{fig:maps1} shows some of the ELMs with the overlap being $0\%, 20\%, 40\%$ respectively, the prior strength being $\alpha = 0.02, 0.06, \dots, 0.24$, and the noise level $p=0.00$. The local maxima corresponding to the correct bi-clusters (either the target bi-cluster or its transposition) are marked with solid red circles; the empty bi-cluster is marked with a gray circle; and the maximal bi-cluster containing the whole data matrix is marked with a solid green circle. These ELMs can be divided into three regimes. \begin{itemize} \item Regime I: the true model is easily learnable; the global maxima correspond to the correct bi-clusters and there are fewer than 6 local minima. \item Regime II: the prior is too strong, the ELM has a dominating minimum which is the maximal bi-cluster. Thus the model is biased and cannot recover the underlying bi-cluster. \item Regime III: the prior is too weak, resulting in too many local minima at similar energy levels. The true model may not be easily learned, although it is possible to obtain approximately correct solutions. \end{itemize} Thus we transfer the table in Figure~\ref{fig:maps1} into a ``difficulty map''. Figure \ref{fig:maps2}(a) shows the difficulty map with three regimes with a noise level $p=0.00$; Figure \ref{fig:maps2}(b) shows the difficulty map with $p=0.02$. Such difficulty maps visualize the effects of various conditions and parameters and thus can be useful in choosing and configuring the biclustering algorithms. \begin{figure*} \center \subfigure[] { \includegraphics[width=\textwidth]{figures/biClusterTrees.png} } \caption{ Energy Landscape Maps for learning two bi-clusters with 0\%,20\%, 40\% overlap and hyperparameter $\alpha$. Red: correct bi-cluster; Grey: empty bi-cluster; Green: maximal bi-cluster. } \label{fig:maps1} \end{figure*} \begin{figure*} \center \subfigure[Noise $p=0.00$]{ \includegraphics[width=0.4 \textwidth]{figures/convergenceRegionsWNoise.png} } \subfigure[Noise $p=0.02$]{ \includegraphics[width=0.4 \textwidth]{figures/convergenceRegion2.png} } \caption{Difficulty map for bi-clustering (a) without noise (b) with noise. Region I: the true model is easily learnable. Region II: the true model cannot be learned. Region III: approximations to the true model may be learned with some difficulty.} \label{fig:maps2} \end{figure*} \section{Conclusion and Discussion} We present a method for computing the energy landscape maps (ELMs) in model spaces for cluster and bi-cluster learning, and thus visualize for the first time the non-convex energy minimization problems in statistical learning. By plotting the ELMs, we have shown how different problem settings, such as separability, levels of supervision, levels of noise, and strength of prior impact the complexity of the energy landscape. We have also compared the behaviors of different learning algorithms in the ELMs. Our study leads to the following problems which are worth exploring in future work. \begin{enumerate} \item If we repeatedly smooth the energy function, adjacent branches in the ELM will gradually be merged, and this produces a series of ELMs representing the coarser structures of the landscape. These ELMs construct the scale space of the landscape. From this scale space ELM, we shall be able to study the difficulty of the underlying learning problem. \item One way to control the scale space of ELM is to design a learning strategy. It starts with lower-dimensional space, simple examples, and proper amount of supervision, and thus the ELM is almost convex. Once the learning algorithm reaches the global minimum of this ELM, we increase the number of hard examples and dimensions and the ELM becomes increasingly complex. Hopefully, the minimum reached in the previous ELM will be close to the global minimum of ELM at the next stage. We have studied a case of such \emph{curriculum learning} on dependency grammars in \cite{Maria_curriculum}. \item The clustering models are defined on Or-And graph structure (here, 'or' means mixture and 'and' means conjunction of dimensions, and the bi-clustering models are defined on And-Or graph. In general, it was shown that many advanced learning problems are defined on hierarchical and compositional graphs, which is summarized as multi-layers of And-Or graphs in~ \cite{Zhu06as}. Studying the ELMs for such models will be more challenging but crucial for many learning tasks of practical importance. \end{enumerate}
\section{Introduction} A key goal of extragalactic astrophysics is the accurate census and measurement of the cosmic star formation rate density (SFRD) across cosmic time. The first measurements of the SFRD \citep{tinsley80a,lilly95a,madau96a} revealed that the density of star formation has decreased tenfold over the last seven billion years since $z\sim1$. More recent measurements \citep*[like those summarized in][]{hopkins06a} show the nature of the SFRD at the highest redshifts, reaching a plateau at $z\sim2$ and steadily declining at earlier times \citep[now extended to $z\sim10$ with very deep near-infrared data, e.g.][]{oesch13a}. However, there is one major caveat to SFRD measurements derived from direct starlight in the ultraviolet, optical and near-infrared, and that is the effect of dust obscuration. Detailed studies of dust attenuation in nearby star-forming regions and star-forming galaxies \citep[e.g.][]{calzetti94a,calzetti01a,meurer99a,overzier11a} have long served as a calibration tool for understanding infrared reprocessed emission in galaxies out to high-redshift. One critical tool has been the empirically observed tight correlation between galaxies' rest-frame ultraviolet (UV) slope, defined as $\beta$, and the ratio of infrared luminosity to UV luminosity at $\approx$1600\,\AA, defined as IRX$\equiv$\,$L_{\rm IR}$/$L_{\rm UV}$. This local relationship between $\beta$ and IRX\ has been widely used as a method to infer dust obscuration, thus total star formation rates, in high-redshift galaxies in the absence of far-infrared data \citep[e.g.][]{bouwens09a}. Thanks to substantial recent development in deep, far-infrared instrumentation \citep*[see the review of ][for a complete discussion of far-infrared datasets used at high-redshift]{casey14a}, the robustness of this IRX--$\beta$\ relationship can now be tested at high-redshifts and high-luminosities. Until recently, the analysis of the rest-frame UV properties of dusty galaxies was limited to small, inhomogeneous samples. Here we combine the strengths of recent large area far-infrared surveys from \mbox{\it Herschel}\ with the extensive 2\,deg$^2$ 30$+$ band UV/optical/near-infrared photometry in the COSMOS field \citep{capak07a,scoville07a,scoville13a} to investigate the rest-frame UV characteristics of large samples of high-$z$ dusty star-forming galaxies (DSFGs). We compare the analysis of the $z>0$ DSFGs to data of local galaxies of all luminosities \citep{gil-de-paz07a,howell10a,u12a} to investigate the underlying physical characteristics of dust attenuation at both low and high-redshifts. Section~\ref{sec:irxbhistory} describes some of the relevant history of the IRX--$\beta$\ relation, and section~\ref{sec:selection} describes our galaxy samples$-$both nearby and at $z>0$ in the COSMOS field$-$as well as presenting basic calculations. Our analysis of DSFGs in the IRX--$\beta$\ plane is presented in Section~\ref{sec:analysis}. We address the possible contamination of high-$z$ LBG dropout studies in Section~\ref{sec:highz}. Our discussion is later presented in Section~\ref{sec:discussion} and conclusions in Section~\ref{sec:conclusions}. Throughout, we assume a Salpeter initial mass function \citep[][although we note conversion to a different IMF, thus different star formation rate, is straightforward]{salpeter55a} and a flat $\Lambda$CDM cosmology \citep{hinshaw09a} with $H_{\rm 0}=71\,$km\,s$^{-1}$\,Mpc$^{-1}$ and $\Omega_{\rm M}=0.27$. \section{The IRX--$\beta$\ Relation To-Date}\label{sec:irxbhistory} The relationship between galaxies' relative dust attenuation, measured as the ratio of IR (8--1000\,$\mu$m) to UV (1600\,\AA) luminosity $L_{\rm IR}$/$L_{\rm UV}$$\equiv$IRX, and their rest-frame ultraviolet color was first studied in a sample of nearby starburst galaxies \citep{meurer95a,meurer99a}. Investigating the ultraviolet emission in nearby galaxies must be done from space, and the first observations to contribute to this area came from the {\it International Ultraviolet Explorer (IUE)} satellite \citep{kinney93a}. While the {\it IUE} played a critical role in laying the groundwork for UV analyses of galaxies, and establishing our understanding of the IRX--$\beta$\ relationship, a key limitation of {\it IUE} data was its small aperture/field of view: 10\arcsec$\times$20\arcsec, typically much smaller than the full spatial extent of nearby galaxies (a few arcminutes across). It should not then be surprising that the {\it IUE} data of \citet{kinney93a} was focused on galaxy cores, which led to an underprediction of galaxies' total UV luminosities. In contrast to the UV, {\it IRAS} 12--100\,$\mu$m\ observations \citep{neugebauer84a} and subsequent far-infrared observations of nearby galaxies \citep[e.g.][]{dunne03a,kawada07a}, have been limited to very large apertures. Without matched apertures in the UV and IR, the original analysis of the IRX--$\beta$\ relation in \citet{meurer99a} was biased by overestimating $IRX$, even though their selection of blue compact dwarfs attempted to circumvent this problem. Furthermore, the {\it IUE} focus on only galaxy cores implied that there could also likely be a UV color bias, with potential underestimation of the global $\beta$ by only pointing towards the blue, UV-bright cores. Thanks to later observations from the {\it Galaxy Evolution Explorer} \citep[{\it GALEX};][]{morrissey07a} which provided wide field-of-view UV photometry for the same nearby galaxies \citep[][]{gil-de-paz07a}, a recent, revised analysis of the \citeauthor{meurer99a} relation finds lower values of $IRX$ and redder UV slopes for the exact same galaxies \citep[see][as well as some discussion in \citealt{overzier11a}]{takeuchi12a}. Besides differences in photometry, it has long been known that galaxies of different types present differently in the IRX--$\beta$\ plane. Young, metal-poor galaxies like the SMC and LMC are redder and less dusty than starbursts, and normal galaxies lie between the young SMC-type galaxies and compact blue starbursts \citep{buat05a,buat10a,seibert05a,cortese06a,boissier07a,boquien09a,boquien12a,munoz-mateos09a,takeuchi10a,hao11a,overzier11a}. Much of these differences are also likely caused by the differences in interal attenuation curve, whether steep, shallow and with or without the 2175\AA\ feature \citep{gordon00a,burgarella05a}. \citet{kong04a} provided a theoretical framework for the interpretation of these differences by parametrizing galaxy types with a birthrate parameter, $b$, defined as the ratio of present to past-averaged star formation rate, whereby starbursts will have a much higher fraction of FUV emission ($\sim$0.153$\mu$m) to NUV emission ($\sim$0.231$\mu$m) from O and B stars. While \citeauthor{kong04a} attribute the difference between normal galaxies and starbursts to a difference in star-formation histories, \citet{seibert05a} and \citet{cortese06a} argue using {\it GALEX}\ data that the differences are likely due to different dust geometries. Of particular interest for this paper is the observation that ultraluminous infrared galaxies (ULIRGs) and related IR-bright galaxy populations lie significantly {\it above} the IRX--$\beta$\ relation, where $IRX$ and $\beta$ have been claimed to be completely uncorrelated \citep{goldader02a,burgarella05a,buat05a,howell10a,takeuchi10a}. \citet{howell10a} interpret sources lying above the IRX--$\beta$\ relation as having an excess of dust, and that the difference in IRX ($\Delta$IRX) from the expected relation represents the extent to which the IR and UV emission is decoupled. \citet{boquien09a} and \citet{munoz-mateos09a} perform detailed studies of nearby galaxies and conclude that both dust geometry and star-formation history have substantial impact on the placement of galaxies on the IRX--$\beta$\ relation. Given the heightened importance of dusty ULIRGs to cosmic star formation at $z\sim1$ and beyond \citep{le-floch05a,caputi07a,casey12b,casey12c}, it seems crucial to understand IRX--$\beta$\ in dusty galaxies as well as unobscured galaxies. To date, this has been unclear. Beyond the nearby Universe, the \citet{meurer99a} IRX--$\beta$\ relation has played a fundamental role in estimating the amount of dust attenuation at high-redshifts, particularly at $z>3$ where direct infrared observations are unavailable \citep{ouchi04a,stanway05a,hathi08a,bouwens09a}. For example, \citet{bouwens09a} use measurements of the rest-frame UV slope $\beta$ of very faint near-IR detected galaxies at $6<z_{\rm phot}<10$ to constrain the dustiness of the early Universe. Because \citeauthor{bouwens09a} find that sources in high-$z$ surveys are significantly bluer than low-$z$ galaxies, they conclude that the dust obscuration plays an insignificant role in galaxy evolution at $z>5$. Given that the nominal IRX--$\beta$\ relation is actually quite uncertain (and based on potentially biased galaxy samples) this constraint on high-$z$ dust obscuration needs independent verification from infrared/submillimeter surveys. Constraining dust obscuration (and star formation rates) at the highest redshifts requires a thorough understanding and calibration of the IRX--$\beta$\ relation beyond the local Universe. Unfortunately, calibrations of the IRX--$\beta$\ relation have been very limited due to longstanding limitations and sensitivity of far-infrared observations. The most thorough analysis of the IRX--$\beta$\ relation at high-$z$ has come from studies of spectroscopically confirmed $z\sim2$ Lyman Break Galaxies \citep[LBGs;][]{reddy06a,reddy09a,reddy10a,reddy12a}. While earlier works (those before about $\sim$2010) largely relied on indirect measurements of galaxies' far-infrared luminosity, \citet{reddy12a} use some of the deepest pointings of the \mbox{\it Herschel Space Observatory}\ in the GOODS fields \citep{elbaz11a} to investigate the direct far-infrared emission in LBGs. Since very few LBGs are directly detected with \mbox{\it Herschel}, they used a stacking analysis to measure the characteristic $L_{\rm IR}$ for $z\sim2$ LBGs and found $\langle L_{\rm IR}({\rm LBG})\rangle \sim$10$^{11}${\rm\,L$_\odot$}; this characteristic luminosity indicates that $\sim$80\%\ of the star formation is obscured in L$_\ast$ galaxies at $z\sim2$. This lines up with expectation from the nominal \citet{meurer99a} IRX--$\beta$\ attenuation curve (in spite of its known problems caused by {\it IUE} aperture limitations and lack of accommodation for `normal' type galaxies). Recently, work on $z\sim4$ LBGs \citep{to14a}, using radio continuum measurements instead of direct far-infrared data, show further agreement with the \citeauthor{meurer99a} relation. Similar to the \citeauthor{reddy12a} results, \citet{heinis13a} explore the dust attenuation law in large samples of UV-selected galaxies at $z\sim1.5$ and find a roughly constant IRX\ ratio over a wide range of UV luminosity explored and an IRX--$\beta$\ relation in line with local `normal' star-forming galaxies. Further stacking results beyond $z>3$ potentially hint at a breakdown in the relation for highly luminous LBGs \citep{lee12a,coppin14a}. While the \citeauthor{reddy12a} and \citeauthor{heinis13a} direct far-infrared studies have shed valuable light on dust attenuation calibrations at high-$z$, they both address the problem using UV-selected galaxy samples, which preferentially might have bluer UV colors and lower IRX\ values than the average galaxy at high-redshift. In contrast, \citet{penner12a} present an analysis of the IRX--$\beta$\ relationship for 24$\mu$m-selected dust-obscured galaxies \citep[DOGs;][]{dey08a}, which have direct detections in the far-infrared from \mbox{\it Herschel}-{\sc pacs} \citep{poglitsch10a}. They find, perhaps not surprisingly, that dustier galaxies have higher $L_{\rm IR}$/$L_{\rm UV}$\ ratios than those selected at UV wavelengths, even at matched rest-frame UV slopes. This is quite similar to earlier work on local ULIRGs \citep{goldader02a,howell10a} which found similarly high $L_{\rm IR}$/$L_{\rm UV}$\ ratios, above expectation from IRX--$\beta$. This difference between `normal' star-forming galaxies and infrared-luminous starbursts is attributed to emergent UV emission not corresponding to the same spatial regions of dust absorption and re-emission \citep{trentham99a,papovich06a,bauer11a}. Indeed, some select studies describing the rest-frame UV and optical properties of submillimeter galaxies \citep[SMGs, selected at 850\,$\mu$m;][]{smail97a} at high-redshift reinforce this result by finding little correspondence between UV luminosity, UV color, and FIR luminosity \citep{frayer00a,smail04a,chapman05a,casey09b,walter12a,fu12a}. This evidence has so far indicated that the nominal IRX--$\beta$\ relation should not be used in dusty galaxies (and vice versa, that any relationship determined for dusty galaxies should not be assumed for a general galaxy population), but what more can we learn? \section{Galaxy Samples and Calculations}\label{sec:selection} We use two sets of galaxies in this analysis; the first is a set of nearby galaxies ($z<0.085$) spanning the characteristic range of galaxy environments in the local volume, with $L_{\rm bol}\sim$10$^{(8-12.5)}$\,{\rm\,L$_\odot$}\ and star formation rates $\approx$0.01--100{\rm\,M$_\odot$\,yr$^{-1}$}. The second galaxy population consists of far-infrared selected star-forming galaxies spanning a redshift range $0<z<5$, with most at $z<3.5$. The selection, data-sets, and description of these samples follow. The calculation of relevant quantities like $L_{\rm UV}$, $L_{\rm IR}$, IRX\ and $\beta$ are performed in a consistent manner between local and $z>0$ samples. \subsection{Nearby Galaxies} Our local dataset combines the 1034 nearby galaxies observed by {\it GALEX}\ included in \citet{gil-de-paz07a}, originally selected from the {\it GALEX}\ Nearby Galaxies Survey (NGS), and the 202 nearby GOALS \citep{armus09a} LIRGs and ULIRGs observed by {\it GALEX}\ \citep{howell10a}, originally selected from the {\it IRAS}\ Revised Bright Galaxy Sample \citep[RBGS;][]{sanders03a}. Note that the {\it GALEX}\ photometry for several of the sources in the \citet{howell10a} are slightly revised in \citet{u12a} to account for highly irregular morphologies associated with galaxy interactions. The IR and UV apertures are matched and taken as total. It is worth noting that all of the original blue compact galaxies used to derive the IRX--$\beta$\ relation in \citet{meurer99a} are included in the \citet{gil-de-paz07a} sample; most of them are significantly redder (higher $\beta$ value) when measured with {\it GALEX}, consistent with other works that have since updated the nominal IRX--$\beta$\ relation \citep{takeuchi10a,takeuchi12a}. \subsection{$0<z<3.5$ Dusty Galaxies} We use a sample of $>$4000 \mbox{\it Herschel}-selected DSFGs in the COSMOS field \citep{scoville07a,scoville13a,capak07a} as our $z>0$ sample, which have full UV-to-radio photometric coverage. \mbox{\it Herschel}\ \citep{pilbratt10a} coverage of COSMOS was carried out with the {\sc pacs} \citep{poglitsch10a} and {\sc spire} \citep{griffin10a} instruments as part of the PEP \citep{lutz11a} and HerMES \citep{oliver12a} surveys. The DSFGs' characteristics and selection ($>$3$\sigma$ in two or more of the five \mbox{\it Herschel}\ bands from 100--500\,$\mu$m) are described in detail in \citet{lee13a}, who present their empirically determined spectral energy distributions from the UV through the far-infrared. The method we use to estimate galaxies' deboosted far-infrared flux densities is described in \citet{roseboom10a} with updates in \citet{roseboom12a}. The sample can largely be regarded as a luminosity-limited sample, as selection is unbiased with respect to far-infrared SED shape (i.e. dust temperature); deboosted flux densities are estimated to be complete above $S_{\rm deboosted}\approx$10\,mJy at 160\,$\mu$m, 250\,$\mu$m, 350\,$\mu$m, and 500\,$\mu$m\ and above $\approx$6\,mJy at 100\,$\mu$m. We refer the reader to \citet{lee13a} for details on sample selection. DSFGs' astrometry and multiwavelength counterparts are identified via a cross-identification method using 24$\mu$m\ and 1.4\,GHz position priors down to $S_{24}=80$\,\uJy\ and $S_{\rm 1.4GHz}=$65\,\uJy\ \citep{roseboom10a,roseboom12a}. This technique is predicated on the notion that high-redshift DSFGs should either be 24$\mu$m\ or 1.4\,GHz {\it detected}; this is an accurate assumption at $z\simlt2$ \citep[with only a 3\%\ failure rate, see][]{magdis11a}, but the sample is largely largely incomplete at higher redshifts where 24$\mu$m\ and 1.4\,GHz surveys are relatively insensitive to detecting $z>2$ galaxies. Of the original 4546 {\it Herschel}-identified galaxies with 24$\mu$m\ or 1.4\,GHz counterparts, only 4165 have reliable photometric redshifts \citep{ilbert09a}: $\approx$92\%\ of the total sample. The accuracy of the photometric redshifts in COSMOS is reported as $\sigma_{\Delta z/(1+z)}=0.012$. It should be noted that the DSFGs without reliable redshifts {\it do} exhibit different optical/near-IR photometry than those with reliable redshifts. Most are undetected in all but a few bands. Without redshifts, it is difficult to characterize this subset of the population with respect to IRX--$\beta$, beyond claiming that they likely have IRX$>$3. Since emission from active galactic nuclei (AGN) has the potential to affect the results of this study$-$by potentially skewing the rest-frame UV colors bluer and boosting the IR luminosities in the mid-IR regime$-$we must remove sources with X-ray detections, strong powerlaws in \mbox{\it Spitzer}\ IRAC and MIPS bands and sources with unexpectedly high radio-to-FIR ratios. Of the 4218 \mbox{\it Herschel}-detected sources in COSMOS, only 5 are directly detected by XMM \citep{brusa10a} and a further 95 are detected by {\it Chandra} \citep{civano12a} suggesting a powerful AGN. They are removed from our analysis. While obscured AGN are less likely to contaminate the rest-frame UV colors of their host galaxies, we also remove 332 additional galaxies suspected of hosting such AGN, evidenced by a strong powerlaw through \mbox{\it Spitzer}\ IRAC and MIPS bands \citep{donley12a}. All COSMOS DSFGs which exhibit very high radio-to-FIR ratios have been identified as AGN through X-ray selection or mid-infrared powerlaw. Of the remaining 4165 DSFGs in our study with photometric redshifts, a total of 432 ($\approx$10\%) have been removed as potential AGN contaminants, leaving 3733 DSFGs for our analysis. Figure~\ref{fig:nz} plots the IR luminosities and redshifts of the COSMOS DSFG sample alongside the local sample for comparison. While 98\%\ of the DSFG sample sits at $z<3.5$, the median redshift for the COSMOS sample is 0.83, in close agreement with the measured peak redshift for {\it Herschel}-bright galaxies \citep{casey12b}. \begin{figure} \centering \includegraphics[width=0.99\columnwidth]{nz.pdf} \caption{IR luminosity against redshift for both the local sample (black points) and the COSMOS $z>0$ sample (blue points). Inset are histograms contrasting their IR luminosities (left inset) and redshifts (right inset). The local sample sits below $z=0.085$ while the median redshift for the COSMOS sample is 0.83, with 60\%\ of the sample sitting at $0.1<z<1$ and 98\%\ below $z$ of 3.5.} \label{fig:nz} \end{figure} \subsection{Calculations}\label{sec:calculations} We fit $L_{\rm IR}$ consistently between the local and $z>0$ samples, from 8 to 1000\,$\mu$m\ and using all available infrared to millimeter data at rest-frame 8--2000\,$\mu$m. For the local sample, this is primarily based on the four {\it IRAS} bands at 12$\mu$m, 25$\mu$m, 60\,$\mu$m\ and 100\,$\mu$m, with some additional data such as {\sc Scuba} \citep{dunne03a} for the brightest nearby (U)LIRGs. The full detailed photometry is described in \citet{u12a}. The COSMOS sample includes the five-band \mbox{\it Herschel}\ 100--500\,$\mu$m\ data, 24--70\,$\mu$m\ \mbox{\it Spitzer}\ data, and when available AzTEC 1.1\,mm data \citep{scott08a,aretxaga11a} as well as {\sc Scuba-2} 450\,$\mu$m\ and 850\,$\mu$m\ data \citep{casey13a,roseboom13a}. We use the far-infrared SED fitting technique described by \citet{casey12a}; this has been shown to produce $L_{\rm IR}$, $T_{\rm dust}$ and $M_{\rm dust}$ estimates that are more accurate (when compared to direct interpolation of data) yet fully consistent with template fitting methods popular in the literature \citep*[e.g.][]{chary01a,dale02a}. \begin{figure*} \centering \includegraphics[width=0.99\columnwidth]{irxbeta_local.pdf} \includegraphics[width=0.99\columnwidth]{irxbeta.pdf} \caption{IRX\ plotted against $\beta$, where $IRX\equiv$$L_{\rm IR}$/$L_{\rm UV}$\ and $\beta$ is the rest-frame UV slope, for local galaxies (left panel) and for $z>0$ galaxies (right panel). In both panels we overlay the original relation from \citet{meurer99a} for blue compact starbursts observed with {\it IUE} and {\it IRAS} (dashed gray line), the local Lyman-break analog (LBA) relation determined in \citet{overzier11a}, which comprises a larger sample (dotted gray line). The revised relation to the \citet{meurer99a} galaxies from \citet{takeuchi12a} includes the aperture effects between {\it IUE} and {\it GALEX}, where the latter correctly accounts for the integrated UV luminosity. On the left panel we plot the local sample of 1034 nearby galaxies described by \citet{gil-de-paz07a} as well as 135 local (U)LIRGs from \citet{howell10a} with $\beta$ and IRX\ re-computed in this paper. This local sample comprises a representative sample of all galaxies in the local volume; all galaxies are color-coded by their star formation rates from $\sim$0.03--100{\rm\,M$_\odot$\,yr$^{-1}$}. Our best-fit attenuation curve, irrespective of galaxy type, is shown in solid black with gray uncertainty envelope (given by Eq.~\ref{eq:irxbeta}). Black square points denote the median UV slope in fixed IRX\ bins. On the right panel we copy our measurement of the local relation and overplot several measurements of IRX--$\beta$\ at high-redshift in galaxy samples that have direct-FIR measurements for $L_{\rm IR}$. A stack of LBGs from \citet{reddy12a} is shown with a navy diamond, split into two $\beta$ bins (navy crosses), and low-luminosity $z\sim2$ ULIRGs (blue diamond). Dust-Obscured Galaxies (DOGs), summarized in \citet{penner12a}, are shown in dark teal, while a comparable ULIRG-type control population is shown in lighter teal. Burnt orange X's denote LBG stacking results in HerMES fields from \citet{heinis13a}. The direct-detection COSMOS DSFGs in this work, spanning redshifts $0<z<5$, mostly at $z<2$, are shown in tan contours with individual outliers overplotted.} \label{fig:irxbeta} \end{figure*} We compute the rest-frame UV luminosity of DSFGs by interpolating their measured photometry to rest-frame 1600\,\AA\ as an AB apparent magnitude, $m_{\rm 1600}$, with \begin{equation} L_{\rm UV} = \frac{4\pi\,D_{\rm L}^2\nu_{\rm 1600}}{(1+z)} \times 10^{-(48.60 + m_{\rm 1600})/2.5}, \label{eq:uv} \end{equation} where $D_{L}$ is the luminosity distance at redshift $z$ and $L_{\rm 1600}$ is given in erg\,s$^{-1}$. At $z\approx0$, UV luminosity is the interpolation of FUV and NUV luminosity to 1600\,\AA\ (which is approximately equal to $L_{\rm FUV}$ given the FUV filter's proximity to rest-frame 1600\,\AA). Note that as the redshifts of the galaxies in the sample increase, different observed bands are representative of the rest-frame UV: the NUV {\it GALEX} 0.231$\mu$m\ filter at $z\simlt0.6$, the Subaru $U$-band 0.346$\mu$m\ filter at $0.6\lower.5ex\hbox{\ltsima} z\simlt1.5$, and the Subaru $B$-band 0.460\,$\mu$m\ filter at $1.5\lower.5ex\hbox{\ltsima} z\simlt2$, and so on. We then compute the rest-frame UV slope, $\beta$, using a powerlaw fit, where $\beta$ is defined by the relationship between flux and wavelength such that $F(\lambda)\propto\lambda^{\beta}$ (where flux is given in units of erg\,s$^{-1}$\,cm$^{-2}$\,\AA$^{-1}$). We calculate $\beta$ and the uncertainty on $\beta$ based on the multiple photometric measurements in the rest-frame UV regime available in COSMOS, namely 1230--3200\,\AA\ \citep[][ this range is chosen to primarily consist of UV radiation from young O and B stars while also being wide enough to make the measurement of $\beta$ feasible using multiple photometric bands]{calzetti94a,meurer99a,calzetti01a}. Depending on redshift, $\beta$ is calculated with 2--5 photometric bands, with an average of 3.3 bands per source. To determine if any systematic bias is introduced in the estimation of $\beta$ as a function of redshift (thus, as a function of the different broad and narrow bands available), we artificially redshift several different galaxy templates \citep*[from][]{bruzual03a}, convolve the templates with the available filter profiles, measure $\beta$ and compare to the intrinsic rest-frame UV slope. While we infer an intrinsic scatter in $\beta$ of up to 0.3 based on filter combination and noise, we infer no systematic variation. Note that since the measurement of $\beta$ is done photometrically, it has the potential to be contaminated by stellar or interstellar absorption features and also enhanced extinction around the rest-frame 2175\,\AA\ dust feature seen in high-metallicity environments like the Milky Way \citep[but is less pronounced in low-metallicity environments like the LMC and SMC;][]{calzetti94a,gordon03a}. This 2175\,\AA\ absorption feature is now known to exist in both low-$z$ galaxies \citep{conroy10a,wild11a} and some high-$z$ galaxies \citep{buat11a,buat12a,kriek13a}; unfortunately, the intrinsic attenuation curve in highly dusty galaxies is more difficult to constrain since fewer rest-frame UV photons escape. We do not think this substantially affects our results if the galaxies we study follow a Calzetti attenuation law, but we emphasize that further work$-$especially on the most extreme starbursts$-$is necessary to establish an empirical constraint on the relationship of attenuation to metallicity. Another potential bias introduced by performing this calculation photometrically, is possible contamination from emission lines, primarily \lya. If a photometric redshift is slightly lower than a galaxy's intrinsic redshift, emission from \lya\ might boost the perceived rest-frame FUV emission indicating a bluer UV slope. To test for this contamination, we compare the best-fit $\beta$ from the 1230--3200\,\AA\ range with a more restrictive 2000--3200\,\AA\ range, which should be substantially less affected by \lya\ contamination for galaxies where photometric redshifts are better than $\Delta z=0.5$. We find no systematic difference between the two different fits to $\beta$ and therefore argue that emission line contamination of $\beta$ in our sample is negligible and/or unlikely. We have corrected for Galactic extinction effects on both UV luminosity and $\beta$ using the dust maps of \citet*{schlegel98a}, with updates provided by \citet{peek10a}; this is of particular importance for the nearby sample, which is distributed across the sky. We use the Milky Way attenuation curve \citep[$A_{\lambda}/A_V$ from][]{gordon03a} and the average extinction to reddening relation at $V-$band of $A_V = 3.1 E(B-V)$ for the diffuse Milky Way \citep{cardelli89a} to determine appropriate correction factors $A_{\lambda}$ for observed bands from the FUV to z band. \section{Analysis}\label{sec:analysis} \subsection{IRX--$\beta$\ for Nearby Galaxies}\label{sec:irxb_local} The left panel of Figure~\ref{fig:irxbeta} shows the IRX--$\beta$\ relation in the local \citet{gil-de-paz07a} and \citet{howell10a} samples after re-measuring $L_{\rm IR}$, $L_{\rm UV}$, and $\beta$ uniformly as described in \S~\ref{sec:calculations}. Galaxies' star formation rates are denoted with color, ranging from blue to red, roughly increasing monotonically with IRX. It is clear that of the three calibrations of this relationship in the local Universe, that of \citet{takeuchi12a} provides the best fit, with both the widely-used \citet{meurer99a} and \citet{overzier11a} fits offset in $\beta$ by $\sim$0.7 towards the blue. Again, this disagreement arises from the corrected, integrated aperture of {\it GALEX}\ between the \citet{meurer99a} and \citet{takeuchi12a} works. While \citet{overzier11a} {\it do} correct for aperture effects, they include a sample of particularly blue Lyman-break Analogs (LBAs) at low redshift which causes the median fit to be substantially bluer than that of \citet{takeuchi12a}. All three prior literature works are blueward of our fit at low IRX\ because of the subset of galaxies used to derive the fit: starbursts. Our inclusion of the \citet{howell10a} LIRG data, and the full \citet{gil-de-paz07a} heterogeneous sample spans a larger dynamic range in star formation rates. If we exclude LIRGs and low SFR systems, we recover a shallower, bluer slope. To determine the best-fit attenuation curve for local galaxies, we first bin up our data in IRX\ intervals of 0.25\,dex. We choose to bin in IRX\ instead of $\beta$, as has been done elsewhere in the literature, to avoid degeneracies at $\beta<-1.5$ (where a wide range of IRX\ values all correspond to the same $\beta$) and at $\beta>0.5$ (where dusty galaxies become non-negligible with high values of IRX\ for the same $\beta$). We note that binning in $\beta$ or IRX\ produces consistent results between $-1.4<\beta<0$ and $1<IRX<50$. Since these data are notably non-Gaussian in $\beta$ at a given IRX, we compute a representative `mode' $\beta$ value by measuring the peak of the $\beta$ histogram for the given IRX\ bin; the black squares on Figure~\ref{fig:irxbeta} are these mode values. Errors on the mode are determined by bootstrapping. We exclude galaxies with IRX$>60$ from the fit, where there is clear deviation, a topic we will return to in subsequent sections. A fit including the data at IRX$>60$ is very shallow and a poor solution to all data above IRX$>1$; we determined the cutoff of IRX$>60$ iteratively by excluding the highest IRX\ bins until the reduced-$\chi^2$ was below 1.5. Note that the galaxies at very high-$\beta$ and low IRX\ are too few to affect the average binning or fit. We determine that \begin{equation} IRX = 1.68\times[10^{0.4[(3.36\pm0.10)+(2.04\pm0.08)\beta]}-1] \label{eq:irxbeta} \end{equation} provides a good fit to the local data, with characteristic spread in $\beta$ of 0.59. This spread represents the standard deviation in the difference of measured $\beta$ values from the expected $\beta$ value (at a given IRX\ and given Eq.~\ref{eq:irxbeta}). Here the rest-frame UV extinction in magnitudes is represented by the quantity A$_{UV}=3.36+2.04\beta$. The factor of 1.68 accounts for the bolometric correction of the original 40--120\,$\mu$m\ FIR studies to total IR luminosities ($L_{\rm IR}$ integrated from 8 to 1000\,$\mu$m). One can verify that 1.68 is an appropriate choice of bolometric correction by generating SEDs with a variety of temperatures \citep[e.g. by using the fitting method of][]{casey12a} and comparing the integral luminosities between the two wavelength ranges. Regardless, the exact value of the bolometric correction only is relevant for readers wishing to compare our best-fit $A_{UV}$ to other works that use similar notation. Very dusty systems with $IRX>60$ are unexpectedly blue. Beyond $IRX>60$, the mode value of UV color $\beta$ hits a wall at $\beta\approx-0.1$. We note that this result is slightly different from the analysis of the local sample presented in \citet{overzier11a}, who only measured galaxies with $IRX\simgt100$ as slight outliers; the difference is primarily due to the redder colors we measure in the low luminosity, low-$IRX$ systems than any discrepancy at the high-$IRX$ end. Also note that, while bluer than expected from the IRX--$\beta$\ relation, these very dusty galaxies are not as blue as some Lyman-Break Galaxies in the early Universe. For the rest of our analysis, unless stated otherwise, whenever we refer to the `local' or `nominal' IRX--$\beta$\ relation, we are explicitly referring to the relationship derived here, represented by Equation~\ref{eq:irxbeta}. \subsection{Context of IRX--$\beta$\ at high-$z$} \begin{figure*} \centering \includegraphics[width=1.99\columnwidth]{irxbgrad.pdf} \caption{Three panels illustrating the IRX--$\beta$\ relation with respect to DSFGs' IR luminosity (left), redshift (middle), and UV luminosity (right), as indicated by point color. The nominal IRX--$\beta$\ relation as given in Equation~\ref{eq:irxbeta} is shown in solid black on each panel for context. The higher a galaxy's IR luminosity or redshift, the farther away that galaxy sits from the IRX--$\beta$\ relation towards bluer rest-frame UV colors, i.e. the steepest gradients in IR luminosity and redshift are orthogonal to the IRX--$\beta$\ relation. On the other hand, the steepest gradient in UV luminosity is parallel to the IRX--$\beta$\ relation. } \label{fig:irxbeta_grad} \end{figure*} The right panel of Figure~\ref{fig:irxbeta} places the local IRX--$\beta$\ relationship in context at higher redshift. The local relations, including the best-fit relation in Equation~\ref{eq:irxbeta} are overplotted with some comparative data-sets from the literature, all of which involve direct far-infrared and direct rest-frame UV measurements (other works using indirect methods have been omitted). We include the 114 LBG stacking result from \citet{reddy12a}, whose measurement sits auspiciously along the \citet{meurer99a} local attenuation curve, and their 12-galaxy, low-luminosity ULIRG comparison sample, which sits at slightly elevated IRX\ relative to \citeauthor{meurer99a} We also include the IR-stacking results on 38000$+$ UV-selected $z\sim1.5$ galaxies from \citet{heinis13a}. Note that the \citeauthor{heinis13a} data agree remarkably well with our locally derived IRX--$\beta$\ relation and not the \citeauthor{reddy12a} stacking results. We attribute this disagreement to the fact that the \citeauthor{heinis13a} sample is not color selected. One puzzling aspect of the \citeauthor{heinis13a} sample is the continuation of the relation to very red colors, $\beta\approx1$. What is different in these galaxies at $z\sim1.5$ that allows for this `extra' reddening? We believe sample selection and binning on $\beta$ rather than $IRX$ (in this case out of necessity since the galaxies are not directly detected in the infrared). The \citeauthor{heinis13a} sample does not include galaxies explicitly {\it selected} by their dust emission. As we found with our local sample, an exclusion of galaxies selected at IR wavelengths skews the IRX--$\beta$\ relation towards redder colors \citep[consistent with what was found by][]{takeuchi12a}. As such, we do not attempt to derive a best-fit $z>0$ IRX--$\beta$\ relation, since it is quite clear from Figure~\ref{fig:irxbeta} (right panel) that most DSFGs do not follow a strict IRX--$\beta$\ relationship. This would only be possible if we had a larger dynamic range in IRX, similar to what is available for the local sample. It is quite clear that sample selection impacts the interpretation of IRX--$\beta$. In relation to the UV-selected samples, the aggregate IRX\ and $\beta$ values presented for dusty galaxies in \citet{penner12a} are notably offset in a similar manner as the local LIRGs and ULIRGs of \citet{howell10a}, emphasizing that dusty galaxies are bluer than the nominal IRX--$\beta$\ relation would predict. Where do the COSMOS DSFGs lie in relation to these other $z>0$ measurements? The tan background contours on the right panel of Figure~\ref{fig:irxbeta} represent the COSMOS DSFGs. They range from being directly on, or even below, the local relation to being above the relation by $\sim2\,$dex. In the next subsection we investigate the difference between DSFGs which lie on the relation and those sitting substantially above. DSFG characteristics in the $IRX-\beta$ plane show strong migration with infrared luminosity, ultraviolet luminosity and redshift. The infrared luminosity and redshift migration goes hand-in-hand, whereby sources at the highest redshifts have the highest IR luminosities, based on our DSFG IR flux density based selection. Indeed, Figure~\ref{fig:irxbeta_grad} shows that contours of equal $L_{\rm IR}$ and $z$ roughly trace one another. On the other hand, contours of constant UV luminosity are orthogonal to those of constant $L_{\rm IR}$. This is expected given our IR sample selection and, importantly, assuming that there is only loose correlation between $L_{\rm UV}$ and $L_{\rm IR}$. From this plot, we can clearly see that galaxies with the highest IR luminosities and redshifts lie farthest away from the local attenuation curves in the IRX--$\beta$\ plane. What does this tell us physically? To understand the underlying physical conditions leading to this $L_{\rm IR}$ or $z$-driven migration of DSFGs in the IRX--$\beta$\ plane, we first have to ensure that this is not caused by sample selection effects. \subsection{Testing for Selection Bias} The COSMOS field where the DSFGs are selected has the benefit of very deep ancillary data from the UV through the near-infrared for use in fitting stellar emission \citep{capak07a}. To test for selection biases in the IRX--$\beta$\ plane, we first complete an analysis of the depth of coverage in every rest-frame UV filter and rest-frame IR filter used to calculate $L_{\rm UV}$, $\beta$ and $L_{\rm IR}$. In the rest-frame UV, we determined completeness as a function of magnitude for all galaxies in the field for each filter, using the photometry reported by \citet{ilbert10a}. In the far-infrared, we know the characteristic depth of {\it Herschel} coverage at 100\,$\mu$m, 160\,$\mu$m, 250\,$\mu$m, 350\,$\mu$m\ and 500\,$\mu$m\ as reported by \citet{lutz11a} and \citet{nguyen10a} and how to apply deboosting factors as appropriate to estimate intrinsic flux densities from raw, corrected for confusion noise \citep{roseboom10a,roseboom12a}. This single characteristic flux limit is primarily blurred by the uncertainty in deboosting factors (and to a much lesser extent, variations in instrument noise). \begin{figure} \begin{center} \includegraphics[width=0.9\columnwidth]{selection_demo0.pdf} \includegraphics[width=0.9\columnwidth]{selection_demo1.pdf} \end{center} \caption{An illustration of the selection effects within our DSFG sample in two redshift slices. Light blue points are individual DSFGs that fall in the given redshift bins, and dark blue square points are the median values of $\beta$ for DSFGs falling in the given IRX\ bins. Tan shaded areas isolate accessible IRX--$\beta$\ space where sources are over 95\%, 90\%\ and 80\%\ likely to be included in our survey. The limit at low IRX\ is caused by the lower limit for direct IR detection combined with an upper limit on $L_{\rm UV}$. The limit at high-IRX, which is a shallow function of $\beta$, depends on the depth of the UV/optical bands in COSMOS. We overplot the local relation from Figure~\ref{fig:irxbeta} in black (gray uncertainty) and simulate the selection functions impact on the local relation (round red points). This demonstrates that the offset towards lower $\beta$ in DSFGs is not driven by selection effects. } \label{fig:selection} \end{figure} To tease out the possible selection biases in the IRX--$\beta$\ plane, we construct a grid across $-3<\beta<3$ and $-2<\log(IRX)<4$, with a binsize of 0.05 in both quantities. Similarly, we construct a parallel grid of plausible IR luminosities and redshifts, which reaches far beyond the range of our observed sample: $6<\log(L_{\rm IR})<14$, with a binsize of 0.05 and $0<z<5$ with $\Delta z$=0.01. For a given $\beta$ and IRX, at a given $L_{\rm IR}$ and $z$ we compute $L_{\rm UV}$ (where $L_{\rm UV}\,=\,L_{\rm IR}/IRX$), rest-frame magnitude $m_{1600}$ at 1600\,\AA\ via Eq~\ref{eq:uv}, and assuming $F\propto\lambda^\beta$, we compute the observed AB-magnitudes in the COSMOS filters that span the rest-frame UV at the given redshift $z$. We then add statistical noise (instrumental and confusion) to the measured magnitudes to be consistent with real data noise, remeasure $\beta$, and determine the probability that the given source would be detectable in our survey given the completeness curves described in the prior paragraph. This test is then repeated across the entire grid of possible $\beta$, IRX, $L_{\rm IR}$ and $z$ values\footnote{As an aside, our tests indicate there is no systematic bias in measuring $\beta$ at low luminosities.}, measuring a probability of inclusion in our DSFG sample for each permutation. To complete this test of selection bias, we must also assign a probability of detection based on redshift and infrared luminosity (two of the four independent parameters in the above test for rest-frame UV detectability). This requires translating our \mbox{\it Herschel}\ flux density limits into luminosity, which requires some assumption on IR SED characteristics, primarily dust temperature. \citet{symeonidis13a} and \citet{lee13a} both present an observed trend of increasing dust temperature with luminosity which is {\it unbiased} with respect to temperature, unlike similarly luminous systems selected at 850\,$\mu$m\ \citep{blain04a,chapman04a,casey09b}. This relation is also observed locally, but shifted towards slightly warmer temperatures, perhaps indicative of size differences in the population, local galaxies being more compact \citep{swinbank13a}. This dust temperature shift at high redshifts is explicitly illustrated in figure~25 of \citet*{casey14a}. We adopt a representative SED shape model to characterize the dust temperature, or SED peak wavelength, in terms of redshift and IR luminosity via \begin{equation} \langle\log\lambda_{\rm peak}\rangle = \eta [\log\!L_{\rm IR} - (\log\!L_{\rm 0} + \gamma\log(1+z))] + \mu. \end{equation} At $z=0$, this equation simplifies to $\langle\log\lambda_{\rm peak}\rangle=\eta(\log\!L_{\rm IR}-\log\!L_{0})+\mu$. Here $\eta=-0.062$ is the slope of the correlation, $\log\!L_{0}=10.60$ sets an arbitrary luminosity zero point at $z=0$, and $\mu=1.99$ is the average $\log\lambda_{\rm peak}$ at that luminosity. The redshift evolution of $L_{0}$ is assumed to take the form $(1+z)^{\gamma}$, as is often used to model the evolution of $L_\ast$ in luminosity functions \citep[e.g.][]{caputi07a,goto10a}, with best-fit $\gamma=2.7$. This function has characteristic scatter around $\log\lambda_{\rm peak}$ of 0.045 (Casey {\rm et\ts al.}\ in prep.). This model gives us a reliable and realistic estimate for the limiting luminosity of our \mbox{\it Herschel}\ data. Note, however, that the temperature-redshift dependence of this model only weakly impacts our conclusions regarding detectability in this paper, and only at the low IRX\ end, and is not a function of the rest-frame UV slope $\beta$. Figure~\ref{fig:selection} offers an illustration of our survey's selection effects in the IRX--$\beta$\ plane in two redshift slices. Areas shaded in tan are likely to be covered by the detection limits of COSMOS data. To answer the question of whether or not our deviation towards bluer colors is driven by selection effects, we model a population of 10$^6$ galaxies which cluster about the local IRX--$\beta$\ curve (from Eq.~\ref{eq:irxbeta}) with measured scatter. After applying our selection limits to that sample and binning in IRX, we are left with the rounded red points in Figure~\ref{fig:selection}, which are significantly more red than the observed samples (shown in blue). This leads us to conclude that selection effects do not drive this observed blue $\beta$ offset in DSFGs. \subsection{Comparison to Literature IR-selected Galaxies} Our result that high-SFR, high-$L_{\rm IR}$ galaxies are much bluer than they would nominally be expected to be\footnote{Where the `expected' $\beta$ for a given IRX\ would be described by our Equation~\ref{eq:irxbeta}.} is consistent with previous analyses on smaller case studies of luminous infrared galaxies, both at low-$z$ \citep{goldader02a,howell10a} and high-$z$. The offset from the IRX--$\beta$\ relation was attributed in \citet{goldader02a} to the significant spatial disassociation of the UV-luminous and IR-luminous portions of each galaxy. \citeauthor{goldader02a} speculated (at a time when very little was known about high-$z$ DSFGs) that local ULIRGs might be unusual in that their physical compactness makes it difficult for much rest-frame UV light to escape (i.e. at a given UV slope $\beta$, a local ULIRG will have a higher $L_{\rm IR}$/$L_{\rm UV}$\ ratio than a normal star-forming galaxy due to its compactness) and at high-$z$, this might not be problematic if DSFGs are more spatially extended, as has often been found to be the case \citep[e.g.][]{ivison98a,ivison10c,ivison11a,hodge12a}. The submillimeter galaxy population \citep[SMGs;][]{smail97a} were, like the local ULIRGs, found to be `bluer' than expected given the incredible $L_{\rm IR}$/$L_{\rm UV}$\ ratios present, \simgt100 \citep{smail04a}. \citet{chapman05a} pointed out that this implied that UV-based star formation rates (SFRs) of SMGs underestimated their total SFRs by factors as large as $\sim$120 even {\it after} the UV SFRs were corrected for extinction \citep[cf.][]{adelberger00a}. \citet{penner12a} also found that dusty galaxies$-$even when they are pre-selected to be incredibly red at optical wavelengths$-$have rest-frame UV characteristics bluer than expected given the local attenuation curve. Furthermore, a comparison control sample of IR-selected galaxies that are not DOGs revealed a similar, yet less extreme result, completely in line with expectation given our results. \begin{figure} \includegraphics[width=0.99\columnwidth]{dbetalir.pdf} \includegraphics[width=0.99\columnwidth]{dbetasfr.pdf} \caption{The deviation from the nominal IRX--$\beta$\ relation, measured as a difference in rest-frame UV slope, $\Delta\beta\equiv \beta_{\rm i} - \beta_{\rm exp}$, where $\beta_{\rm exp}$ relates to IRX\ via Equation~\ref{eq:irxbeta}, against IR luminosity (top) and total star formation rate (bottom). Here we use SFR$_{\rm IR}$+SFR$_{\rm UV}$ as a proxy for total star formation rate, which above $\lower.5ex\hbox{\gtsima}$10\,{\rm\,M$_\odot$\,yr$^{-1}$}\ has 90\%\ of its energy output in the IR (see top axis of bottom plot). The local samples are shown as a shaded gray backdrop and overplotted black median values, while the $z>0$ COSMOS DSFGs are contoured in tan with median values in dark red. The median redshift of the sample at a given SFR is shown in the bottom inset ranging from $0<z<3$. Literature studies with published $L_{\rm IR}$ values are overplotted with the same symbols as in Figure~\ref{fig:irxbeta} (right). In both diagrams, we see a strong break at $L_{\rm IR}\approx10^{11.5}${\rm\,L$_\odot$}\ or SFR$\approx40\,${\rm\,M$_\odot$}\,yr$^{-1}$, above which galaxies of all epochs are bluer than expectation.} \label{fig:dbeta} \end{figure} \begin{figure} \includegraphics[width=0.99\columnwidth]{dbetaz.pdf} \caption{Same as Figure~\ref{fig:dbeta} but split into $\Delta z=0.2$ redshift bins to probe a possible underlying redshift evolution distinct from the overall luminosity-driven IRX--$\beta$\ deviation. The weighted mean $\Delta\beta$ value for a given redshift bin and luminosity bin is shown with a bootstrap-estimated uncertainty. The only significant redshift evolution which is observed is seen between $0.6<z<1.4$ and $2-8\times10^{11}${\rm\,L$_\odot$}\ (central box), whereby galaxies at lower redshift are redder by $\Delta\beta=1$. Galaxies at $z<0.6$ and $L_{\rm IR}<2\times10^{11}${\rm\,L$_\odot$}\ are too few in number to measure evolution, while evolution is simply not seen in the higher luminosity bins above $z>1.4$. The inset plot shows the change in $\Delta\beta$ with redshift within this interval for the three luminosity bins at 2.4$\times$10$^{11}${\rm\,L$_\odot$}\ (dotted), 4.2$\times$10$^{11}${\rm\,L$_\odot$}\ (solid), and 7.5$\times$10$^{11}${\rm\,L$_\odot$}\ (dashed). } \label{fig:dbetaz} \end{figure} \subsection{Deviation from the IRX--$\beta$\ relation} From Figure~\ref{fig:irxbeta_grad} we gather that the deviation from the nominal IRX--$\beta$\ relation is either a result of IR luminosity or redshift. Because local ULIRGs also deviate from IRX--$\beta$, we attribute the correlation of galaxies' deviation to their IR luminosities. Figure~\ref{fig:dbeta} investigates this deviation as a function of $L_{\rm IR}$. Due to the degeneracy of IRX\ at low $\beta$, we measure the deviation from the nominal IRX--$\beta$\ relation (that given by Eq.~\ref{eq:irxbeta}) as the difference in UV-slope, or $\Delta\beta$. \citet{howell10a} present a similar plot in terms of $\Delta$IRX for the local sample (black/gray shaded background on our plot). Although there is large scatter in $\Delta\beta$, owing to the complex star formation histories in a heterogeneous galaxy population, both the local and $z>0$ samples share a characteristic `break' luminosity, above which galaxies systematically deviate from IRX--$\beta$\ towards bluer UV slopes. The strength of the deviation increases with increasing luminosity. The `break' luminosity for the local sample appears to sit at $\approx$10$^{11-11.5}${\rm\,L$_\odot$}\ while the break in the COSMOS DSFG sample lies clearly at $\approx10^{11.5}${\rm\,L$_\odot$}. Following \citet{kong04a} and some discussion presented in \citet{reddy06a}, this deviation at high $L_{\rm IR}$ is anticipated. Galaxies with more intense, more recent star formation will be intrinsically bluer for a fixed dust attenuation, because the underlying emission at UV wavelengths is dominated by a higher proportion of young O stars contributing to the stellar continuum emission at 1216--1600\,\AA. In the dusty star-forming environments of ULIRGs where attenuation is substantial, $L_{\rm IR}$ can be directly mapped to the total SFR. The break in the $L_{\rm IR}-\Delta\beta$ plot correlates with increasing star formation rate. The bottom panel of Figure~\ref{fig:dbeta} investigates $\Delta\beta$ as a direct function of SFR$_{\rm total}$, or SFR$_{\rm IR}+$SFR$_{\rm UV}$. To emphasize the relative contributions of IR and UV to SFR$_{\rm total}$, the top axis of the bottom panel indicates the fractional output of star formation in the infrared. Similar to the break IR luminosity quoted above, the break in SFR is seen at $\approx30-50${\rm\,M$_\odot$\,yr$^{-1}$}\ in both local and $z>0$ samples \citep[assuming a Salpeter IMF;][]{salpeter55a}. Although the deviation from IRX--$\beta$\ is clearly systematic above a given $L_{\rm IR}$, the huge scatter of $\sigma_{\Delta\beta}=1$ leads us to ask whether or not there is also any underlying redshift evolution. Figure~\ref{fig:dbetaz} breaks up the $L_{\rm IR}-\Delta\beta$ plot into redshift bins with $\Delta z=0.2$, where the tracks in redshift are representative of the weighted mean and uncertainties are bootstrapped. Slight differences between median $\Delta\beta$ exist with redshift and are seen most prominently in the half decade of $L_{\rm IR}\approx10^{11.5-12}${\rm\,L$_\odot$}. The two luminosity bins at $4\times10^{11}$ and $7.5\times10^{11}${\rm\,L$_\odot$}\ show substantial evolution between $z=0.6$ and $z=1.4$. Over the corresponding cosmic time, the median UV color in galaxies of equal luminosity shifts by $\Delta\beta=1$, i.e. it is substantially redder at lower redshifts. At higher luminosities (and also higher redshifts), no significant differences are detected between epochs. \begin{deluxetable*}{l@{ }c@{ }c@{ }c@{ }c@{ }c@{ }c@{ }c} \tablecolumns{5} \tablecaption{DSFG Contaminants to high-$z$ LBG searches} \tablehead{ \colhead{Target} & \colhead{Selection Criteria} & \multicolumn{5}{c}{\underline{\sc DSFG Contaminant Characteristics}} & \colhead{\%} \\ \colhead{redshift} & \colhead{} & \colhead{$N_{\rm contam}$} & \colhead{$\rho_{\rm contam}$} & \colhead{Av. app.} & \colhead{$z_{\rm phot}$} & \colhead{Av. implied} & \colhead{LBG} \\ \colhead{} & \colhead{} & \colhead{} & \colhead{[arcmin$^{-2}$]} & \colhead{mag} & \colhead{} & \colhead{$\beta$} & \colhead{{\sc contam.}} \\} \startdata \multicolumn{8}{c}{{\sc Optical-Only LBG Dropout Selection}} \\ \hline $z\sim2.5$ & {\footnotesize [$U_{\rm 300}-B_{\rm 450}>1.1$] $\wedge$ [$B_{\rm 450}-I_{\rm 814}<1.5$] $\wedge$} & 15 & 2.4$\times$10$^{-3}$ & ($B$)26.7 & 2.7 & $-0.37\pm0.13$ & 0.01\%\ \\ & {\footnotesize [$U_{\rm 300}-B_{\rm 450}>0.66$($B_{\rm 450}-I_{\rm 814}$)$+1.1$] $\wedge$ [$U>27$] } & & & & \\ $z\sim4$ & {\footnotesize [$B_{\rm 435}-V_{\rm 606}>1.1$] $\wedge$ [$V_{\rm 606}-z_{\rm 850}<1.6$] $\wedge$} & 35 & 5.5$\times$10$^{-3}$ & ($V$)26.3 & 2.9 & $-0.30\pm0.18$ & 0.37\%\ \\ & {\footnotesize [$B_{\rm 435}-V_{\rm 606}>$($V_{\rm 606}-z_{\rm 850}$)$+1.1$] $\wedge$ [$UB>27$]} & & & & \\ $z\sim5$ & {\footnotesize [$V_{\rm 606}-i_{\rm 775}>1.2$] $\wedge$ [$i_{\rm 775}-z_{\rm 850}<0.6$] $\wedge$} & 26 & 4.1$\times$10$^{-3}$ & ($i$)26.2 & 2.0 & $0.0\pm0.2$ & 3.2\%\ \\ & {\footnotesize [($V_{\rm 606}-i_{\rm 775}>0.9$($i_{\rm 775}-z_{\rm 850}$))$\vee$($V_{\rm 606}-i_{\rm 775}>2$)] } & & & & \\ & {\footnotesize $\wedge$ ($UBV>27$)} & & & & \\ $z\sim6$ & {\footnotesize [$i_{\rm 775}-z_{\rm 850}>1.3$] $\wedge$ [[$z_{\rm 850}-J_{\rm 110}<0.6$] $\vee$} & 7 & 1.4$\times$10$^{-3}$ & ($z$)26.1 & 2.0 & $-0.4\pm0.2$ & 8.7\%\ \\ & {\footnotesize [$i_{\rm 775}-z_{\rm 850}>2/3$($z_{\rm 850}-J_{\rm 110}$)$+1.02$]] $\wedge$ [$UBVi>27$]} & & & \\ \hline \multicolumn{8}{c}{{\sc Optical and Near-IR LBG Dropout Selection}} \\ \hline $z\sim4$ & {\footnotesize [$B_{\rm 435}-V_{\rm 606}>1$] $\wedge$ [$i_{\rm 775}-J_{\rm 125}<1$] $\wedge$} & 6 & 9.4$\times$10$^{-4}$ & ($V$)26.3 & 2.9 & $-0.30\pm0.18$ & $<$0.01\%\ \\ & {\footnotesize [$B_{\rm 435}-V_{\rm 606}>1.6$($i_{\rm 775}-J_{\rm 125}$)$+1$] $\wedge$ [$UB>27$]} & & & & \\ $z\sim5$ & {\footnotesize [$V_{\rm 606}-i_{\rm 775}>1.2$] $\wedge$ [$z_{\rm 850}-H_{\rm 160}<1.3$] $\wedge$} & 2 & 3.2$\times$10$^{-4}$ & ($i$)26.2 & 2.0 & $0.0\pm0.2$ & 0.01\%\ \\ & {\footnotesize [$V_{\rm 606}-i_{\rm 775}>0.8$($z_{\rm 850}-H_{\rm 160}$)$+1.2$] } & & & & \\ & {\footnotesize $\wedge$ ($UBV>27$)} & & & & \\ $z\sim6$ & {\footnotesize [$i_{\rm 775}-z_{\rm 850}>1.0$] $\wedge$ [$Y_{\rm 105}-H_{\rm 160}<0.45$] $\wedge$} & 0 & $<$1.6$\times$10$^{-4}$ & ($z$)26.1 & 2.0 & $-0.4\pm0.2$ & $<$0.04\%\ \\ & {\footnotesize [$i_{\rm 775}-z_{\rm 850}>0.777$($Y_{\rm 105}-H_{\rm 160}$)$+1.0$]] $\wedge$ [$UBVi>27$]} & & & & \\ $z\sim7$ & {\footnotesize [$z_{\rm 850}-Y_{105}>0.7$] $\wedge$ [$J_{\rm 125}-H_{\rm 160}<0.45$] $\wedge$ } & 3 & 4.7$\times$10$^{-4}$ & ($Y$)24.9 & 1.9 & $-1.0\pm0.2$ & 0.40\%\ \\ & {\footnotesize [$z_{\rm 850}-Y_{\rm 105}>0.8$($J_{\rm 125}-H_{\rm 160}$)$+0.7$] $\wedge$ [$UBViz>27$]} & & & & \\ $z\sim8$ & {\footnotesize [$Y_{\rm 105}-J_{\rm 125}>0.45$] $\wedge$ [$J_{\rm 125}-H_{\rm 160}<0.5$] $\wedge$ } & 1 & 1.6$\times$10$^{-4}$ & ($J$)24.1 & 4.3 & $-1.11\pm0.14$ & 0.34\%\ \\ & {\footnotesize [$Y_{\rm 105}-J_{\rm 125}>0.75$($J_{\rm 125}-H_{\rm 160}$)$+0.525$]} $\wedge$ [$UBVizY>27$] & & & & \\ $z\sim10$ & {\footnotesize [$J_{\rm 125}-H_{\rm 160}>1.2$] $\wedge$ [$H_{\rm 160}-$[$3.6$]$<1.4$ $\vee$} & 0 & 3.2$\times$10$^{-4}$ & ($H$)23.0 & 2.8 & $-0.2\pm0.4$ & $<$2.5\%\ \\ & {\footnotesize $S/N([3.6])<2$] $\wedge $[$UBVizYJ>27$]} & & & & \\ \enddata \label{tab:contam} \tablecomments{ The above selection criteria are outlined explicitly in \citet{bouwens09a}, for optical-only selection, and \citet{bouwens14a}, for optical and near-IR selection, yet the selection method is very similar to other high-$z$ dropout selection techniques outlined in \citet{bunker03a}, \citet{giavalisco04b}, \citet{beckwith06a}, \citet{stanway07a}, \citet{oesch10a}, and \citet{bouwens11a}. Magnitudes are interpolated from the observed COSMOS broad-, intermediate- and narrow-band imaging to the given selection filters and the effective area of this search is performed in the 1.76\,deg$^{2}$ area of UltraVISTA in COSMOS ($\lower.5ex\hbox{\gtsima}$6 times the area used in the deep dropout searches). In addition to the stated color selection criteria, we use a S/N cutoff in $UBVizYJ$ bands as appropriate. The characteristics of the DSFGs satisfying the selection criteria are given: their average apparent magnitudes, implied $\beta$ if at the target redshift, their median photometric redshift \citep[from the UltraVISTA near-IR selected catalog;][]{mccracken12a} and their density on the sky. The last column gives the percentage of LBG candidates, selected at the given target redshift, that are in fact DSFGs at lower redshifts. We use the stated number of candidates identified per square arcminute as stated in \citet{bouwens09a} and \citet{bouwens14a}. } \end{deluxetable*} Two plausible explanations for this observed `reddening' of matched-$L_{\rm IR}$ galaxies seen between $0.6<z<1.4$ are (a) an increasing metallicity of galaxies towards lower redshifts, or (b) different star formation histories present in galaxies at $z=1.4$ vs.\ $z=0.6$. In the former case, it should be noted that some studies investigating the rest-frame UV continuum properties of optically-selected galaxy populations show evidence for intrinsically bluer colors in lower metallicity systems \citep{alavi14a,castellano14a}, perhaps owing to a more top-heavy IMF in lower metallicity galaxies \citep[e.g.][]{marks12a}. Note also that \citet{marks12a} argue that a more top-heavy IMF will manifest in environments with dense molecular clouds like starbursts, resulting in bluer rest-frame UV slopes. The second explanation for the observed reddening with epoch is a different star formation history between LIRGs at $z=0.6$ and $z=1.4$. Following \citet{kong04a}, we note that galaxies with more recent burst histories and younger stellar populations are expected to have bluer rest-frame UV slopes. This suggests, in fact, that our $z=1.4$ DSFGs are likely {\it younger} and {\it burstier} than their $z=0.6$ analogues, a notion which might seem contradictory to recent works suggesting that LIRGs and ULIRGs at high-$z$ are {\it less} likely to be burst-driven than their local counterparts, based on the galaxy main-sequence \citep[e.g.][]{noeske07a,elbaz11a,rodighiero11a,nordon12a}. Besides redshift evolution, the large scatter in color in each luminosity bin could also be partly due to differences in the intrinsic dust attenuation curves, where flatter curves indicate larger attenuations (IRX) for the same color \citep[$\beta$][]{gordon00a,burgarella05a}, or viewing angle effects. Studying the morphological characteristics of the bluest and reddest sources of a given luminosity or star formation rate to probe inclination, bulge-to-disk ratio, and interactions will be a necessary and worthy follow-up study yet is beyond the scope of this paper. It is worth highlighting that, although we offer many plausible physical explanations here to describe the deviation (or possible bluewards $z$-evolution), we primarily attribute the bulk shift off IRX--$\beta$\ to dust geometry effects. As we will later discuss in more detail in \S~\ref{sec:discussion}, we know that nearby LIRGs and ULIRGs are primarily enshrouded in a thick cocoon of dust, where the IR emission is high and the UV photons are few. The UV light that does escape does so in a patchy pattern, leaking out in bright concentrations. These small openings dominate the UV luminosity, therefore the total UV color of the galaxy, while the dust enshrowded component (which is spatially dis-associated) dominates $L_{\rm IR}$. We return to this in our discussion. Note that the various other measurements of the IRX--$\beta$\ relation at $z>0$ from LBGs and DOGs, align with the observed deviation we measure for DSFGs at their respective IR luminosities, within uncertainty. Could this explain some of the discrepancies between, e.g. the \citet{reddy12a} result$-$LBGs that lie blueward of our nominal IRX--$\beta$\ fit$-$the \citet{heinis13a} results$-$UV selected galaxies lying very close to our nominal IRX--$\beta$\ fit$-$and the \citet{penner12a} results$-$where dust-selected galaxies are notably bluer than our nominal IRX--$\beta$\ fit? We think yes, that it may be understood as a function of these' galaxies total star formation rates or IR luminosities. So while earlier we thought it auspicious that the $z\sim2$ \citeauthor{reddy12a} LBGs aligned perfectly with the \citet{meurer99a} IRX--$\beta$\ relation (a relation with known biases and measurement discrepancies), here we can attribute that alignment to two different biases which together are skewed blueward of the nominal IRX--$\beta$\ relation for a heterogeneous, normal galaxy population. It is also not surprising to see the much bluer colors in the \citet{penner12a} DOGs given their high IR luminosities. \section{How do DSFGs impact high-$z$ galaxy searches?}\label{sec:highz} The search for the highest redshift galaxies, at $z>4$, is predicated on the Lyman Break dropout technique \citep{steidel96a}. DSFGs are usually thought to be too intrinsically red and too rare to impact high-$z$ LBG searches, but our results hint that a subset of DSFGs might satisfy LBG selection criteria, either because they are bluer than anticipated or, alternatively, their rest-frame optical emission lines might contribute substantially to broad-band photometry and be mistaken for a Lyman break at higher redshift \citep[cf.][]{bouwens11b}. Although the LBG dropout selection technique has some advantages, in providing an easily repeatable selection, it is potentially prone to more contamination than high-$z$ photometric redshift techniques, which are less biased against LBGs with intrinsically redder colors \citep{finkelstein10a,finkelstein12a}. This section makes use of the extensive, deep optical and near-IR data in the COSMOS field to explore the extent to which directly-detected \mbox{\it Herschel}\ DSFGs might contaminate high-$z$ LBG `dropout' searches. LBG `dropout' selection at high-$z$ is defined usually in terms of {\it Hubble Space Telescope} broad band filters. We use the 30+ band photometry in the COSMOS field \citep[most recently complete with deep near-IR imaging from UltraVISTA covering 1.76\,deg$^2$;][]{mccracken12a} to interpolate and predict magnitudes in these {\it HST} filters and then apply the various LBG selection criteria to our DSFGs, as outlined in Table~\ref{tab:contam}. We use the LBG dropout selections that are outlined in \citet{bouwens09a}, for optical data only, and \citet{bouwens14a}, for optical and near-infrared data. In addition to the color cuts, we apply a magnitude cutoff of 27 [AB] in bands shortward of the \lya\ break. This prevents against obvious, lower redshift bright contaminants and is preferred to a strict S/N cutoff since the COSMOS data are of varying depths compared to the deep {\it HST} imaging used to identify $z\simgt4$ LBGs. The contaminating fraction of DSFGs in LBG searches is then computed by comparing their surface densities to those of LBG candidates of comparable magnitudes \citep[as presented in][]{bouwens09a,bouwens14a}. Table~\ref{tab:contam} summarizes our results. We find that at low redshifts, $z\simlt4$, the contamination of LBG searches with DSFGs is negligible ($\ll1\%$), since LBGs at that epoch are extremely common relative to DSFGs. The DSFGs which do satisfy the LBG color cuts at these epochs do seem to sit at the appropriate redshifts and are the bluest of DSFGs, but again, are negligible in number when compared to the much more common LBG. Without near-infrared selection criteria, dusty galaxies can easily pass LBG selection at $z\sim5-6$ and comprise a significant fraction of contaminants, as much as $\approx$9\%\ at $z\sim6$. However, thanks to the recent deep surveys from WFC3 in the near-infrared, most of those contaminants are thrown out, with rates as low as 0.1\%. Figure~\ref{fig:highzcontam} shows an average template for DSFGs that contaminate these high-$z$ LBG dropout searches. To construct the template, we first build a median template for contaminants in each LBG selection, corresponding to different redshifts, and then shift them into the same rest-frame wavelength grid: that which corresponds to the assumed redshift of selected LBGs. The median template for all of the LBG selection is then shown in blue\footnote{Available for download now at herschel.uci.edu/cmcasey/research.html under `Tools' and in the future available on the PI's research website under the same header.}, clearly showing the corresponding break that is mistaken for the Lyman-$\alpha$ break. Note that this template is not meant to be physical as it is a coaddition of DSFGs' observed SEDs at a variety of redshifts, and as such, the break observed is caused by a number of different physical processes, e.g. bright [O{\sc ii}] emission, bright {\rm\,H$\alpha$}, or a 4000\,\AA\ or Balmer break. One primary difference seen between the expected SED of an LBG \citep[see the composite in Fig~\ref{fig:highzcontam}][]{shapley03a,calzetti94a} and our contaminant template is that the rest-frame optical emission of LBGs is bluer than in DSFG contaminants. Although this method of checking for dusty `contaminants' in UV galaxy searches has caveats, most notably mismatch photometric depths of the surveys and the likely incompleteness of DSFG samples beyond $z>2$, it serves as a worthy reality check for high-$z$ galaxy search campaigns. The increasing rate of contamination in the highest redshift bins is due to both the degeneracies of colors over a relatively short rest-frame wavelength span and the diminishing surface density of higher-$z$ LBGs. Overall, contamination rates of $<2-3\%$ from dusty starbursts bode well for high-$z$ searches. Besides providing an essential measurement of the Universe's star formation rate density at very early times, high-$z$ LBG searches have led us to infer the dust content of the earliest galaxies using the IRX--$\beta$\ relation \citep{bouwens09a}. Because the distribution of rest-frame UV slopes of LBGs is bluer at earlier times, \citeauthor{bouwens09a} argued that these high-$z$ LBGs contain very little dust. Despite the degeneracy between blue dust-less and dusty galaxies, our results$-$low contamination rates from DSFGs in LBG searches$-$corroborate the \citeauthor{bouwens09a} conclusions, that the majority of high-$z$ LBGs contain relatively little dust, by demonstrating that very few high-$z$ LBGs will be directly detected with {\it Herschel} or similar submm instruments. Of course deep submillimeter follow-up (e.g. from ALMA) is needed to infer the dust content in individual high-$z$ LBGs \citep[e.g.][]{chapman09b}, in particular to constrain their SFRs and contribution to the total star formation rate density. This is because our results have only ruled out the tip of the iceberg: it is quite possible that blue LBGs in the early Universe are much dustier than their UV slopes may infer yet still too faint to be detected in existing \mbox{\it Herschel}\ data. Furthermore, the existence of extreme, dusty galaxies at the same epochs should raise concern for our understanding of the integrated star formation rate density \citep[SFRD, e.g.][]{hopkins06a} at these early times. The star formation rate in one DSFG can often exceed the SFR in individual LBGs by factors of 50--100. While here we have determined that a very low fraction of LBGs will be directly-detected by {\it Herschel}, we note that: (a) {\it Herschel} is far less sensitive to detecting $z>2$ starbursts than longer-wavelength submm surveys \citep*[at $\approx$1\,mm, see figure 7 of][]{casey14a}; and (b) very few $z\sim4-6$ DSFGs will be selected as LBGs. Of the 39 COSMOS DSFGs with photometric redshifts above 4, only 20\%\ satisfy LBG selection criteria. Furthermore, we know that high-$z$ extreme DSFGs like HFLS3 at $z=6.34$ {\it do} exist \citep{riechers13a,cooray14a}, yet the difficulty in obtaining spectroscopic identification has made the assessment of their volume density quite challenging \citep{dowell14a}. While our comparison to LBGs illustrates that there is little concern for significant contamination, we emphasize that the $z>4$ SFRD is still highly uncertain without concrete constraints from direct far-infrared measurements. \begin{figure} \centering \includegraphics[width=0.99\columnwidth]{highzcontam.pdf} \caption{Average template SED for the DSFG population that contaminates LBG dropout searches, given in rest-frame wavelength at the intended target redshift. In gray, we show the individual photometric points for DSFG contaminants which satisfy each of the different redshift LBG cuts given in Table~\ref{tab:contam}. The blue curve gives the average contaminant SED for the entire DSFG contaminating population, averaged over all selection redshifts (in most cases, the DSFG redshifts are dissimilar to the target redshifts). This contaminant template illustrates the need for rest-frame optical constraints on LBG SEDs that are anticipated to be slightly bluer than most DSFG contaminants. Note that very deep coverage shortward of the Lyman break does not guarantee high-redshift identification, since most DSFGs also drop out of the bluest optical filters. The DSFGs' photometry is observed and no normalization to a flux scale is done.} \label{fig:highzcontam} \end{figure} \section{Discussion}\label{sec:discussion} Our results indicate that galaxies with particularly high star formation rates, most of which are measured from their output in the far-infrared, are bluer than expected given the nominal dust attenuation curve for low-luminosity, `normal' star-forming galaxies. What underlying physical processes could be responsible for these bluer UV colors? The first possible explanation is that the rest-frame UV emission and far-infrared emission are physically dis-associated, or spatially distinct \citep{goldader02a,chapman04b}. Assuming that the systems are still physically bound, this might be due to a recent catastrophic event, like a merger-driven burst. Can geometric effects alone explain the systematically bluer rest-frame UV slopes in DSFGs? If a system consists of two dominant components, one with $L_{\rm UV}\approx L_{\rm IR}$ (unobscured) and one with $L_{\rm IR}\gg L_{\rm UV}$ (obscured), both of which follow the IRX--$\beta$\ relation, the IRX\ value assumed for the whole system will be taken from the obscured component, since there the IR luminosity dominates, and the $\beta$ value from the unobscured component, which dominates all UV light; the result is an unresolved, blue DSFG with high IRX. But take another system, which has its IR luminosity and UV luminosity well distributed in, e.g., a disk (for a simple test, if you were to split it into two components, as before, both would have roughly evenly matched $L_{\rm IR}$ and $L_{\rm UV}$). If the individual regions of this homogenized system all followed IRX--$\beta$, then the total would as well. Only dramatic spatial variations in the distribution of $L_{\rm UV}$ and $L_{\rm IR}$ can lift a system substantially above IRX--$\beta$. Figure~\ref{fig:geometry} illustrates how the relative dustiness (IRX\ values) between the two hypothetical, nearby components impacts how `blue' the aggregate sum of the two will be relative to IRX--$\beta$. We constructed this figure by simulating $10^6$ pairs of components, where each indidivdual component sits on the IRX--$\beta$\ relation. To explain a sample of very blue DSFGs systematically offset from IRX--$\beta$\ using this geometrical argument, the vast majority of DSFGs would need to be comprised of two radially different components (where IRX\ of the first would be a factor of $\lower.5ex\hbox{\gtsima}$100--1000 times larger than the second). In other words, the distributions of IRX\ and $\beta$ over the spatial extent of a DSFG would need to be bimodal. Indeed, this is effectively what we see locally (e.g. Arp\,220, Mrk\,273, IRAS\,19254$-$7245). The subcomponents of local ULIRGs which are infrared-bright produce very little UV light and the patches which are UV-bright are blue and see little IR contribution. Aside from physical dis-association, alternate explanations also exist, although with less observational backing amongst local analogs. Nevertheless, we discuss them here since the effect we observe could be due in part to these multiple factors. As described in \S~\ref{sec:analysis}, \citet{kong04a} suggest that IRX--$\beta$\ should not, in fact, be a fixed relation for all galaxies, but rather it should vary for galaxies with different star formation histories. The more recent and intense the star-formation, the bluer the galaxy (for a fixed $L_{\rm IR}$/$L_{\rm UV}$\ ratio). While overall this interpretation explains Figure~\ref{fig:dbeta} qualitatively, we note that such a pronounced `break' above $L_{\rm IR}\approx10^{11.5}${\rm\,L$_\odot$}\ might not be expected. Whether or not it should be expected depends significantly on the given model assumptions for the effective absorption curve, and whether or not stars irradiate just their birth clouds or the ambient ISM \citep*[suggested to transition at a stellar age of 10\,Myr;][]{charlot00a}, as well as the distribution of the stars with respect to the ISM, the lifetime of burst episodes (as short as 10\,Myr to as long as 300\,Myr), and the characteristics of the underlying long-term star formation. \citet{kong04a} provide a sensible model construct for quiescent and normal galaxies, but do not offer an explanation for sources at low $\beta$ and high IRX. Those galaxies could be caught during a very short duration burst, where a substantial fraction of the emitted UV radiation is trapped in stars' birth clouds before they have time to migrate out. \begin{figure} \includegraphics[width=0.99\columnwidth]{geometry.pdf} \caption{Can spatially disassociated UV- and IR-bright components explain the bluer rest-frame UV colors of DSFGs? Here we explore how two nearby galaxy components of different obscurations (IRX\ values) might be perceived on the IRX--$\beta$\ plane if the two components are not resolved from one another. We simulate 10$^6$ hypothetical pairs where all individual components lie on the IRX--$\beta$\ relation (dashed black line) with observed local scatter $\pm$0.59 in $\beta$. Only pairs with significant differences between component IRX\ values will be measured as having bluer-than-expected colors due to spatial disassociation. This simplified simulation is effectively modeling the relative bimodality of a galaxy's dust obscuration. We overplot all DSFGs within the narrow $L_{\rm IR}$ ranging corresponding to our simulation in light blue (median values are dark blue squares) for comparison. This figure suggests that most 10$^{12}${\rm\,L$_\odot$}\ DSFGs are bimodal, or have spatially disassociated components which differ in IRX\ by factors of $\sim$300. } \label{fig:geometry} \end{figure} In addition to having a much bluer intrinsic UV slope, DSFGs might have enhanced $L_{\rm IR}$/$L_{\rm UV}$\ ratios due to the short timescale of recent star formation. This might be due to UV radiation being trapped in stars' birth clouds, or more globally, could be explained by the geometry of dust and stars in the galaxy. In a well-mixed ISM with a more spheroidal distribution, the surface-to-volume ratio is lower, and a galaxy with higher SFR will have a higher $L_{\rm IR}$/$L_{\rm UV}$\ ratio, assuming $L_{\rm IR}$ emanates from the whole volume while $L_{\rm UV}$ emerges only from regions close to the surface. On the contrary, lower luminosity, continuous-SFR disk galaxies have lower optical depth for dust attenuation and greater surface area-to-volume ratios, implying (a) $L_{\rm IR}$ and $L_{\rm UV}$ emanate from the whole volume, and (b) the underlying UV continuum is redder than in extreme burst galaxies. Whether or not the magnitude of the shift towards bluer $\beta$ (or higher IRX) for DSFGs is in line with expectation requires more detailed model investigation, which is beyond the scope of this paper. Nevertheless, our results support the notion that DSFGs are likely inconsistent with prolonged, constant star formation histories of disk galaxies, as might be suggested by works favoring the galaxy main-sequence view of DSFGs at high-$z$. In the context of the galaxy main-sequence, readers might be curious how our results would present themselves if, instead of a $L_{\rm IR}$- or SFR-dependent break in $\Delta\beta$, we investigated the break as a function of galaxy stellar mass or specific star formation rate. We have intentionally avoided the use of stellar mass estimates in this paper, primarily due to their reliance on an assumed star formation history. Infrared-luminous galaxies' star formation histories are particularly difficult to determine and have large uncertainties \citep{hainline11a,michaowski12a,michaowski14a}. This in itself is a topic of ongoing debate in the literature \citep*[see more in \S~5 of][]{casey14a}. We hope that follow-up studies that do investigate IRX--$\beta$\ with stellar mass will carefully disentangle various star formation history {\it a priori} assumptions from any conclusions used to interpret galaxy evolution. Delving deeper into the underlying physics behind bluer UV slopes, aside from our simple and favored geometric explanation, the possibility was raised in \S~\ref{sec:analysis} that metallicity and a possible top-heavy IMFs might also contribute to a higher proportion of O starlight, and thus bluer slopes. A top-heavy IMF might be plausible in the dense star-forming regions in ultraluminous galaxies, e.g. DSFGs, consistent with dense star clusters in the Milky Way like Arches and Westerlund~1 \citep{marks12a}. Furthermore, our observation of some evolution in DSFG UV slopes of matched IR luminosity in the range $0.6<z<1.4$ would be consistent with the idea that higher redshift systems are lower metallicity systems. Note that, if we assume that DSFGs at higher redshifts should be less bursty \citep[as many papers suggest despite direct evidence to the contrary, e.g.][]{engel10a,ivison12a}, we might expect a shift in the {\it opposite} direction, towards redder UV slopes, or no shift at all. This is because main sequence galaxies are proposed to have steady state, approximately constant star formation rates for most of their lifetime; even at very high SFRs ($\lower.5ex\hbox{\gtsima}$100{\rm\,M$_\odot$\,yr$^{-1}$}), the expected underlying (unextincted) UV continuum of steady state star formation will be redder than starburst galaxies, which have a disproportionately large contribution of light from O and B stars above the less luminous and older A stars. Again, our results are inconsistent with the suggestion that $>10^{11.5}${\rm\,L$_\odot$}\ DSFGs at high-$z$ are dominated by steady state, secular, disky star formation. \section{Conclusions}\label{sec:conclusions} This paper has investigated the relationship between the $L_{\rm IR}$/$L_{\rm UV}$\ ratio (probing the relative `dustiness' of galaxies) to their rest-frame ultraviolet continuum slope, $\beta$, particularly as it pertains to infrared-selected galaxies. By comparing a sample of $\approx$1200 nearby galaxies ($z<0.085$) spanning star formation rates 0.03--300{\rm\,M$_\odot$\,yr$^{-1}$}\ to a large sample of $\approx$4000 IR-selected galaxies spanning photometric redshifts $0<z<5$ in the COSMOS field, we have arrived at the following conclusions. \begin{enumerate} \item We derive a much redder IRX--$\beta$\ relationship for local galaxies than was presented in some of the original works on the topic \citep[e.g.][]{meurer99a} and attribute the difference to (a) differences between {\it IUE} and {\it GALEX}\ aperture limitations (the latter provides a more representative characteristic estimate for $\beta$); and (b) the fact that we derive the relation for a more heterogeneous population, not limited to galaxies selected as blue, compact starbursts. Our derived relation is given in Equation~\ref{eq:irxbeta}. This is roughly consistent with prior works, which corrected the original relation for aperture differences \citep{takeuchi12a}. \item We find that, at both low and high-redshift, DSFGs with high SFRs above $\approx50${\rm\,M$_\odot$\,yr$^{-1}$} deviate from this IRX--$\beta$\ relation towards bluer colors, where the offset grows with increasing IR luminosity and star formation rate. The deviation towards bluer colors, measured as $\Delta\beta$ is seen both in local and high-$z$ samples above a `break' IR luminosity of $\approx10^{11-11.5}${\rm\,L$_\odot$}\ and increases with increasing luminosity such that at $L_{\rm IR}=10^{13}${\rm\,L$_\odot$}, galaxies are on average $\Delta\beta=-2$ bluer than expected from the nominal IRX--$\beta$\ relation fit in Eq~\ref{eq:irxbeta}. This offset towards bluer colors is shown not to be caused by sample selection effects. \item Subtle redshift evolution is detected in the narrow luminosity regime $4-7.5\times10^{11}${\rm\,L$_\odot$}\ for $0.6<z<1.4$, where galaxies of matched IR luminosity are on average $\Delta\beta=-1$ bluer at $z=1.4$ than at $z=0.6$. No redshift evolution is detected above $z>1.4$ (at representatively higher IR luminosities, corresponding to detection limits at that epoch). More extensive samples of equal luminosity over a wider range of epochs is needed to verify this perceived evolution. \item We attribute the bluer colors in dusty galaxies to more recent, rapid episodes of star formation \citep[following the `birthrate' parameter model offered by][]{kong04a}, where more IR luminous galaxies have a more prominent population of young, O and B stars (contributing to the rest-frame far-UV emission) than galaxies of more modest star formation rates. Not only are they intrinsically bluer, DSFGs likely have higher IRX\ values caused by lower emergent UV-to-IR luminosity ratios and mixed, patchy geometry. This is consistent with the idea that star formation in DSFGs at high redshift is dominated by burst activity rather than steady state, gradual disk growth. \item With deep multi-band UV/optical and near-IR data in hand, we investigated the rates at which DSFGs contaminate high-$z$ LBG dropout searches. Due to the relatively low sky density of DSFGs, we find very low contamination rates at $z\simlt7$ when both deep optical and near-infrared data exist. Contamination rates increase at higher redshifts, up to 5.1\%\ at $z\sim10$, where there is less information on the rest-frame optical and more potential low-redshift interlopers (likewise, we see high contamination, 8.9\%, at $z\sim6$ if observed near-infrared bands are not used for LBG selection). Overall, DSFG contamination rates $<10$\%\ bode well for LBG searches; however, we caution that this does not imply that $z>4$ SFRD estimates from LBGs alone are sufficient for understanding star formation at early times. \end{enumerate} While this work has shed light on the issue of rest-frame ultraviolet emission in the dustiest galaxies, it is clear that much more work is necessary to understand the underlying physics driving these bluer rest-frame UV slopes in dusty galaxies. Geometry, morphological effects, and galaxy interactions need to be constrained on individual systems. Constraining these galaxies' star formation histories via detailed SED fitting is crucial to isolate the dominant physical mechanisms in high-$z$ star formation. The coming years will provide crucial insight into the issue of extinction in extreme galaxies, by studying its links to gas dynamics, star formation rate, galaxy morphology, gas supply, dust reservoir, metallicity, and differences between emission line and continuum extinction. This paper has only provided a broad context from which necessary follow-up studies of detailed systems will reveal the nature of dust attenuation in extreme environments. \acknowledgements COSMOS is based on observations with the NASA/ESA {\it Hubble Space Telescope}, obtained at the Space Telescope Science Institute, which is operated by AURA Inc, under NASA contract NAS 5-26555; also based on data collected at: the Subaru Telescope, which is operated by the National Astronomical Observatory of Japan; the XMM-Newton, an ESA science mission with instruments and contributions directly funded by ESA Member States and NASA; the European Southern Observatory, Chile; Kitt Peak National Observatory, Cerro Tololo Inter-American Observatory, and the National Optical Astronomy Observatory, which are operated by the Association of Universities for Research in Astronomy, Inc. (AURA) under cooperative agreement with the National Science Foundation; the National Radio Astronomy Observatory which is a facility of the National Science Foundation operated under cooperative agreement by Associated Universities, Inc; and the Canada-France-Hawaii Telescope operated by the National Research Council of Canada, the Centre National de la Recherche Scientifique de France and the University of Hawaii. PACS has been developed by a consortium of institutes led by MPE (Germany) and including UVIE (Austria); KU Leuven, CSL, IMEC (Belgium); CEA, LAM (France); MPIA (Germany); INAF-IFSI/OAA/OAP/OAT, LENS, SISSA (Italy); IAC (Spain). This development has been supported by the funding agencies BMVIT (Austria), ESA-PRODEX (Belgium), CEA/CNES (France), DLR (Germany), ASI/INAF (Italy), and CICYT/MCYT (Spain). SPIRE has been developed by a consortium of institutes led by Cardiff Univ. (UK) and including: Univ. Lethbridge (Canada); NAOC (China); CEA, LAM (France); IFSI, Univ. Padua (Italy); IAC (Spain); Stockholm Observatory (Sweden); Imperial College London, RAL, UCL-MSSL, UKATC, Univ. Sussex (UK); and Caltech, JPL, HNSC, Univ. Colorado (USA). This development has been supported by national funding agencies: CSA (Canada); NAOC (China); CEA, CNES, CNRS (France); ASI (Italy); MCINN (Spain); SNSB (Sweden); STFC, UKSA (UK); and NASA (USA). This research has made use of data from the HerMES project (http://hermes.sussex.ac.uk/). HerMES is a \mbox{\it Herschel}\ Key Programme utilizing Guaranteed Time from the SPIRE instrument team, ESAC scientists and a mission scientist. The data presented in this paper will be released through the HerMES Database in Marseille, HeDaM (http://hedam.oamp.fr/HerMES/). We would also like to recognize the use of the Glue Visualization tool (www.glueviz.org) in the initial analysis of this data-set. CMC acknowledges support from a McCue Fellowship through the University of California, Irvine's Center for Cosmology. AC acknowledge support from NSF AST-1313319. RJI acknowledges support from ERC AdG, COSMICISM. TTT has been supported by the Grant-in-Aid for the Scientific Research Fund (24111707) commissioned by the Ministry of Education, Culture, Sports, Science and Technology (MEXT) of Japan. TTT has also been partially supported from Strategic Young Researches Overseas Visits Program for Accelerating Brain Circulation from the MEXT. We are grateful to the anonymous referee for productive comments which improved the paper. We would also like to thank Naveen Reddy, Brian Siana, Mara Salvato and Douglas Scott for helpful conversations during the manuscript's preparation.
\section{Introduction} There has recently been much progress on the exact solution of graph partitioning problems, see for instance the survey~\cite{MeyerhenkeSanders13}, as well as on heuristics for graphs with special structure like for instance road networks. In particular, Delling et al.~\cite{DellingGoldbergRazenshteynWerneck11,DellingGoldbergRazenshteynWerneck12,DellingWerneck12} investigate new combinatorial lower bounds (based on approximations of maximum flow-minimum cut bounds) for the unweighted problem, and perform computational studies within a parallel branch-and-bound framework~\cite{BudiuDellingWerneck11}. Armbruster et al. study linear and semidefinite programming approaches~\cite{ArmbrusterFugenschuhHelmbergMartin12}, and present a sequential computational study. Improved flow-based bounds were given in~\cite{Sensen01}, also with a computational study. This note investigates another, simple, combinatorial lower-bound approach which applies to both the weighted and unweighted graph (bi)partitioning problems. The basic lower bound was originally proposed in the early 90ties~\cite{Traff91:or,Traff94:ppl} with some later improvements~\cite{Traff96:slb}. We present proofs and further improvements, and implementations within the parallel task-scheduling framework Pheet\footnote{The framework with the implementations described in this note can be downloaded from \url{www.pheet.org}} which has been extensively described in~\cite{Wimmer14:diss}. The motivation for this bound is the belief that weaker, but more easily computable bounds may be preferable for parallel branch-and-bound over stronger but hard-to-compute bounds in order to keep a large number of processing units (threads, processes, cores, processors, \ldots) busy throughout the solution of the given partitioning problems. This was observed in~\cite{Traff94:ppl}; and~\cite{BudiuDellingWerneck11,DellingGoldbergRazenshteynWerneck12} give similar motivations for their bounds. \section{The graph partitioning problem} Given a weighted, undirected graph $G=(V,E)$ with vertices (or nodes, used synonymously) $V$ and edges $E$ with arbitrary (real or integer) edge weights $w(u,v), (u,v)\in E$, the \emph{graph bipartitioning problem} is to find a partition (\emph{cut}) of $V$ into two subsets $V_0$ and $V_1$ of given sizes $|V_0|=s_0$ and $|V_1|=s_1$ with $s_0+s_1=n$ and $s_0>0, s_1>0$ having minimum cut weight $w(V_0,V_1)$ over all such partitions. The weight of a cut is defined by extension of the weight function as \begin{eqnarray*} w(V_0,V_1) & = & \sum_{\{(u,v)\in E | u\in V_0,v\in V_1\}}w(u,v) \end{eqnarray*} for any two disjoint subsets $V_0\subset V$ and $V_1\subset V$. The graph partitioning problem is NP-hard, see e.g.~\cite{GareyJohnson79,GareyJohnsonStockmeyer76}. The natural (and relevant) generalization of the problem to partitioning $V$ into $k, k>2$ subsets $V_i$ with predefined sizes $|V_i|=s_i$ (or with predefined total vertex costs) is not discussed here, but many of the observations carry over to the $k$-partitioning problem also. \section{Lower and upper bounds} We solve the graph partitioning problem using \emph{branch-and-bound}, a standard, search based method~\cite{PapadimitriouSteiglitz82} which is presumably well-suited to parallel implementation, see, e.g.~\cite{CrainicLeCunRoucairol06,GendronCrainic94,Talbi06}. The essential components of a branch-and-bound algorithm are the notions of \emph{subproblem}, \emph{completion}, \emph{lower bound}, and \emph{branching rule}. The lower bound provides for any subproblem a bound on the cut value of any completion of the subproblem. As soon as the lower bound for a subproblem is larger than or equal to some current, best feasible solution (or \emph{upper bound}) the subproblem can be discarded from further consideration since it can never lead to a better solution. Any partition of the vertex set $V$ into a pair of subsets $(V_0,V_1)$ that fulfills $|V_i|=s_i, i=0,1$ is a \emph{feasible solution} to the graph partitioning problem. A \emph{subproblem} is a pair $(U_0,U_1)$ of disjoint subsets of $V$ with with $|U_i|\leq s_i,i=0,1$ representing a partial assignment of vertices to either of the two subsets. A \emph{completion} of a subproblem $(U_0,U_1)$ is a feasible solution $(V_0,V_1)$ with $U_i\subseteq V_i,i=0,1$. Vertices of $G$ in either of $U_i$ are said to be \emph{fixed}, otherwise \emph{free}. The set of free nodes is thus $F=V\setminus (U_0 \cup U_1)$. The \emph{branching rule} selects a free node $v\in F$ and creates two new subproblems by extending either of the sets $U_i$ with $v$, such that $(U_0\cup\{v\},U_1)$ and $(U_0,U_1\cup\{v\})$ will be the two new subproblems to be considered. The branch-and-bound process starts from an empty subproblem $(\emptyset,\emptyset)$, respectively, if $n$ is even, from a subproblem $(\{u\},\emptyset)$ for some node $u$ in order to avoid generating symmetric solutions. Let $n=|V|$ and $m=|E|$. We assume that $G$ has no self-loops $(u,u)$; such edges never contribute to a cut anyway. We also assume that edges $(u,v)\in E$ have non-negative costs. For the implementation, we let $V=\{0,\ldots,n-1\}$. We represent a subproblem $(U_0,U_1)$ by two bitmaps $B_i$ of $n$ bits; bit $u$ of $B_i$ is set iff $u\in U_i, i=0,1$. Furthermore, we also maintain a bitmap for the free vertices, and in addition an array of free vertices with $f=|F|=n-|U_0|-|U_1|$ being the number of free vertices. The weighted input graph $G$ is represented by an array of adjacency arrays. Note that each edge $(u,v)\in E$ is present in the adjacency arrays of both node $u$ and of node $v$. We also need for each edge $(u,v)$ in the $i$th position of the adjacency array of $u$ the position $j$ of $u$ in the adjacency array of $v$. Finally, we store the adjacency arrays in sorted, non-decreasing weight order. The lower bounds are based on the following simple observation. Let $(V_0,V_1)$ be a completion of a subproblem $(U_0,U_1)$. It holds that \begin{eqnarray*} w(V_0,V_1) & = & w(U_0,U_1) + w(V_0\setminus U_0,U_1) + w(U_0,V_1\setminus U_1) + w(V_0\setminus U_0,V_1\setminus U_1) \end{eqnarray*} A lower bound for a subproblem $(U_0,U_1)$ is therefore given by the cut between already assigned vertices in $U_0$ and $U_1$, $w(U_0,U_1)$, plus a lower bound on the term $w(V_0\setminus U_0,U_1)+w(U_0,V_1\setminus U_1)$, and finally a lower bound on the term $w(V_0\setminus U_0,V_1\setminus U_1)$. The latter two contributions can be treated independently. \subsection{The lower bound: basic bound and rebalancing} \label{sec:basicrebalance} Let $v\in F$ be a free node in the subproblem $(U_0,U_1)$. Any completion $(V_0,V_1)$ will have a contribution to the cut value from $v$ of at least $\min(w(v,U_0),w(v,U_1))$, no matter whether $v$ is eventually in $V_0$ or $V_1$. Namely, if $v$ is in $V_0$ all edges from $v$ to nodes in $U_1$ will contribute to the cut, and similarly if $v$ is in $V_1$. Thus, a trivial lower bound for the term $w(V_0\setminus U_0,U_1) + w(U_0,V_1\setminus U_1)$ is \begin{eqnarray*} B(U_0,U_1) & = & \sum_{v\in F}\min(w(v,U_0),w(v,U_1)) \end{eqnarray*} Computing this bound from scratch takes $O(n+m)$ steps. If we maintain for each (free) vertex $v$ the two values $D_i[v] = w(v,U_i)$ for the cost of assigning $v$ to subset $V_{i-1}$, the lower bound can be computed as $\sum_{v\in F}\min(D_0[v],D_1[v])$ in $O(f)$ time steps where $f=|F|$ is the number of free nodes. When branching on node $v$ and $v$ is put into $V_i$, all values $D_{i}[u]$ where $(v,u)\in E$ need to be increased by $w(v,u)$. This can be done in $O(\deg(v))$ steps. The bound $B(U_0,U_1)$ does not take the cardinality constraints on completions of $(U_0,U_1)$ into account. If, for instance, $w(v,U_1)<w(v,U_0)$ for a large number of nodes, then $B(U_0,U_1)$ may count too many vertices as having been assigned to subset $V_0$, and a stronger bound could be obtained by counting some of these vertices as assigned to $V_1$. Define $\delta(v) = w(v,U_1)-w(v,U_0)=D_1[v]-D_0[v]$ as the \emph{potential free weight increase} of node $v$. If $\delta(v)>0$ vertex $v$ would tend to be assigned to subset $V_1$, and there is a penalty of $\delta(v)$ of assigning $v$ to $V_0$ instead; if $\delta(v)<0$ the lower bound $B(U_0,U_1)$ would count $v$ as assigned to $V_0$, and there would be a penalty of $-\delta(v)$ of instead assigning $v$ to $V_1$. Penalties are the amounts of which the lower bound might be increased when the cardinality of the sets $V_0\setminus U_0$ and $V_1\setminus U_1$ are taken into account. Let $\delta_i, 0\leq i<f$ be the potential free weight increases in sorted order, $\delta_i\leq \delta_{i+1}$ for $0\leq i<f-1$. Then the lower bound can be strengthened by a \emph{rebalancing contribution} \begin{eqnarray*} R(U_0,U_1) & = & \sum_{i=0}^{f_0-1}\max(0,\delta_i)+\sum_{i=f_0}^{f_1-1}\max(0,-\delta_i) \end{eqnarray*} where $f_i=s_i-|U_i|, i=0,1$. This basic rebalancing bound was first presented in~\cite{Traff91:or,Traff94:ppl}. Computing the rebalancing contribution seems to require sorting of the $\delta(v), v\in F$ values and can be done easily in $O(f\log f)$ steps. Our implementation computes the rebalancing bound in this fashion. Maintaining the $\delta(v)$ values in a priority queue does not improve complexity, since up to $f$ values have to be considered in order, and each extract min operation takes logarithmic time. However, if the $\delta(v)$ values are maintained in sorted order, recomputation and sorting is necessary only for $\deg(v)$ nodes when branching on vertex $v$. The full array of $f$ values can be reestablished by merging. The complexity of the rebalancing steps is hereby reduced to $O(f+\deg(v)\log\deg(v)))$ which is $O(n^2+m\log n)$ for the whole a series of at most $n$ branching steps. \begin{proposition} \label{prop:rebalancing} For any given subproblem $(U_0,U_1)$ it holds that \begin{eqnarray*} B(U_0,U_1)+R(U_0,U_1) & \leq & w(V_0\setminus U_0,U_1)+w(U_0,V_1\setminus U_1) \end{eqnarray*} for any completion $(V_0,V_1)$. The bound is \emph{tight}: there is a completion $(V_0,V_1)$ such that $B(U_0,U_1)+R(U_0,U_1) = w(V_0\setminus U_0,V_1)+w(U_0,V_1\setminus U_1)$. \end{proposition} \begin{proof} As argued above, $B(U_0,U_1)$ is a lower bound on $w(V_0\setminus U_0,V_1)+w(U_0,V_1\setminus U_1)$ in any completion $(V_0,V_1)$: Each assigned vertex will contribute a weight of either $w(v,U_1)$ or $w(U_0,v)$. The crucial part is the rebalancing step. Let $(V_0,V_1)$ be a completion that minimizes $w(V_0\setminus U_0,U_0)+w(U_0,V_1\setminus U_1)$. Pick any two nodes $u\in V_0\setminus U_0$ and $v\in V_1\setminus U_1$. It must hold that $w(u,U_1)+w(v,U_0) \leq w(u,U_0)+w(v,U_1)$ since otherwise the weight $w(V_0\setminus U_0,U_1)+w(U_0,V_1\setminus U_1)$ could be reduced by swapping $u$ and $v$. This implies that $\delta(u)\leq\delta(v)$. Let $f_i=|V_i\setminus U_i|, i=0,1$. We now prove by induction on $\min(f_0,f_1)$ that \begin{eqnarray*} B(U_0,U_1)+R(U_0,U_1) & = & w(V_0\setminus U_0,U_1)+w(U_0,V_1\setminus U_1) \end{eqnarray*} for such a completion. Assume first that either $f_0=0$ or $f_1=0$. If $f_0=0$, all free vertices are assigned to $V_1$, and for each $v$ it holds that $w(v,U_0) = \min(w(v,U_0),w(v,U_1))+\max(0,-\delta(v))$; namely, if $w(v,U_0)>w(v,U_1)$ then $w(v,U_0) = w(v,U_1)-(w(v,U_1)-w(v,U_0))=w(v,U_1)-\delta(v)$. If instead $f_1=0$, then it holds that $w(v,U_1)=\min(w(v,U_0),w(v,U_1))+\max(0,\delta(v))$. Therefore, in either case \begin{eqnarray*} w(V_0\setminus U_0,U_1)+w(U_0,V_1\setminus U_1) & = & \sum_{v\in F}\min(w(v,U_0),w(v,U_1)) + \\ & & \sum_{i=0}^{f_0-1}\max(0,\delta_i)+\sum_{i=f_0}^{f_1-1}\max(0,-\delta_i) \\ & = & B(U_0,U_1)+R(U_0,U_1) \end{eqnarray*} Now assume that $\min(f_0,f_1)>0$. Choose a vertex $u\in V_1\setminus U_0$ that maximizes $\delta(u)$ over all such $u$, and a vertex $v\in V_1\setminus U_1$ that minimizes $\delta(v)$ over all such $v$. Recall that $\delta(u)\leq\delta(v)$. The contribution of $u$ to $w(V_0\setminus U_0,U_1)+w(U_0,V_1\setminus U_1)$ is $w(u,U_1)$, which, if $\delta(u)\geq 0$ can be written as $\min(w(u,U_0),w(u,U_1))+\delta(u) = w(u,U_1)$ since $w(u,U_0)\leq w(u,U_1)$. The contribution of $u\in V_0\setminus U_0$ is therefore $\min(w(u,U_0),w(u,U_1))+\max(0,\delta(u))$. Likewise, the contribution of $v$ to $w(V_0\setminus U_0,U_1)+w(U_0,V_1\setminus U_1)$ is $w(v,U_0)$ which, by a similar case analysis, can be written as $\min(w(v,U_0),w(v,U_1))+\max(0,-\delta(v))$. We can now remove $u$ and $v$ from the set of free edges. The resulting completion $(V_0\setminus\{u\},V_1\setminus\{v\})$ minimizes $w(V_0\setminus\{u\}\setminus U_0,U_1)+w(U_0,V_1\setminus\{v\}\setminus U_1)$, and $\delta(u')\leq \delta(v')$ for all $u'\in V_0\setminus\{u\}\setminus U_0$ and $v'\in V_1\setminus\{v\}\setminus U_1$. By the induction hypothesis \begin{eqnarray*} w(V_0\setminus\{u\}\setminus U_0,U_1)+w(U_0,V_1\setminus\{v\}\setminus U_1) & = & \sum_{v\in F\setminus\{u,v\}}\min(w(v,U_0),w(v,U_1)) + \\ & & \sum_{i=0}^{f'_0-1}\max(0,\delta_i)+\sum_{i=f'_0}^{f'_1-1}\max(0,-\delta_i) \end{eqnarray*} where $f'_i=f_i-1$ is the size of the subsets with $u$ and $v$ removed. Adding in the contribution from $u$ and $v$ establishes the lower bound claim. \end{proof} To represent a subproblem, our implementation uses $O(n)$ for the bitmaps of fixed and free nodes, the array of free nodes, and the $D_i$ values for the free vertices. \subsection{The lower bound: high-degree unassigned vertices} \label{sec:highdegree} The lower bound counts edges between fixed and free vertices in the subproblem $(U_0,U_1)$ but is oblivious to contributions from edges between free nodes. However, some of these edges inevitably contribute to a lower bound on completions of $(U_0,U_1)$. Let wlog $U_0$ be the subset with the largest number of nodes still to be assigned, that is assume that $f_0\geq f_1$. Consider the graph $G'=(F,E')$ induced by the set of free nodes $F$, and let $v\in F$. If the degree of $v$ in $G'$ is larger than $f_0-1$ then there will be at least $\deg'(v)-f_0+1$ edges out of $v$ in any cut of $F$ into subsets of size $f_0$ and $f_1$. Here, $\deg'(v)$ denotes the degree of $v$ in $G'$. The smallest weight such edges are a lower bound for the contribution of $v$. Let $T_i(v) = \sum_{j=0}^{\max(0,\deg'(v)-f_i+1)}w_j(v)$, where $w_j$ is the $j$th smallest weight of an edge in $G'$ adjacent to $v$. Summing these contributions over all free nodes and dividing by two since both vertices of a cut edge may have a lower bound contribution gives \begin{eqnarray*} T(U_0,U_1) & = & \sum_{v\in F}T_0(v)/2 \end{eqnarray*} For integer edge weights, $\ceiling{\sum_{v\in F}T_0(v)/2}$ is still a lower bound, as will follow from the argument below. These observations were first made in~\cite{Traff96:slb}. Also here rebalancing can be applied. If there are more than $f_0$ free nodes of high degree, the lower bound counts too many nodes as becoming assigned to $V_0$. Let $\delta'(v) = T_1(v)-T_0(v)$ be the \emph{penalty} of assigning $v$ to the larger subset of size $f_0$; if $\delta'(v)>0$ there is a gain of assigning $v$ instead to the subset of size $f_1$ (note that for all $v$, $\delta'(v)\geq 0$). Again, let $\delta'_i$ be the penalties in sorted order. Then the rebalancing contribution is \begin{eqnarray*} R'(U_0,U_1) & = & \sum_{i=f_0}^{f_1-1}\delta'_{i-f_0}/2 \end{eqnarray*} Note that rebalancing gives a contribution only if the number of high-degree vertices is larger than $f_0$. \begin{proposition} \label{prop:highdegree} For any given subproblem $(U_0,U_1)$ it holds that \begin{eqnarray*} T(U_0,U_1)+R'(U_0,U_1) & \leq & w(V_0\setminus U_0,V_1\setminus U_1) \end{eqnarray*} for any completion $(V_0,V_1)$. \end{proposition} \begin{proof} We prove that $T(U_0,U_1)$ is a lower bound for the partitioning problem on the graph $G'$ induced by $F$. Let $(u,v)$ be an edge in some cut $(W_0,W_1)$ of $F$ fulfilling the constraints $|W_i|=f_i, i=0,1$. The proof is by induction on the number of cut edges $(u,v)$ where either $u$ or $v$ is adjacent to a high-degree vertex with degree larger than $f_0-1$. Pick one such cut edge. Let $u$ be a high degree vertex, and assume that $(u,v)$ is the $i$th smallest edge adjacent to $u$ in the cut. Note that there must be at least $\deg'(u)-f_0+1$ edges adjacent to $u$ in any cut since $u$ is a high-degree vertex and $G'$ has no self-loops. If $i<\deg(u)-f_0$, the lower bound has a contribution from the $i$th smallest edge $(u,v')$ of weight $w(u,v')$ with $w(u,v')\leq w(u,v)$ (note that $v'$ may be $v$). Similarly if $v$ is a high-degree vertex. The contribution to the lower bound for cut edge $(u,v)$ is $w(u,v')/2+w(v,u')/2$, which is at most $w(u,v)$, since $w(u,v')+w(u',v)\leq 2w(u,v)$. If $v$ is not a high-degree vertex, there is no contribution from $v$, but it still holds that $w(u,v')/2\leq w(u,v)$. The cut edge $(u,v)$ can therefore be removed from the cut, and the edges $(u,v')$ and $(v,u')$ (if present) from $G'$, and the claim now follows by induction. Proof that rebalancing still yields a lower bound can be done by induction, similarly to the proof of Proposition~\ref{prop:rebalancing}. \end{proof} \begin{figure} \centering \begin{tikzpicture}[scale=0.75] \node[dot](a) {}; \node[dot,right=2cm of a] (b){}; \node[dot,below=2cm of b] (c){}; \node[dot,left= 2cm of c] (d){}; \node[dot] at ($(d)!0.25!(a)$)(f){}; \node[dot] at ($(d)!0.5!(a)$)(g){}; \node[dot] at ($(d)!0.75!(a)$)(h){}; \node[dot] at ($(b)!0.5!(c)$)(k){}; \node[dot] at ($(b)!0.75!(c)$)(l){}; \node[ell,rotate=90] at ($(b)!0.5!(c)$) {}; \node[ell,rotate=90] at ($(d)!0.5!(a)$) {}; \draw[line] (a) -- (b); \draw[dashline] (a) -- (c); \draw[dashline] (a) -- (k); \draw[dashline] (a) -- (l); \draw[dashline] (b) -- (f); \draw[dashline] (b) -- (g); \draw[dashline] (b) -- (h); \node[above] at ($(a)!0.5!(b)$) {$3$}; \node[above] at ($(a)!0.9!(k)$) {$2$}; \node[above] at ($(f)!0.1!(b)$) {$2$}; \node[left=3mm] at (a) {$u$}; \node[right=3mm] at (b) {$v$}; \node[right=3mm] at (k) {$v'$}; \node[left=3mm] at (f) {$u'$}; \end{tikzpicture} \caption{The base case for the high-degree lower bound proof. The cut edge $(u,v)$ for two high-degree nodes $u$ and $v$ is shown as a heavy line, the smaller lower bound edges $(u,v')$ and $(v,u')$ as dotted lines; $u'$ and $v'$ may or may not be high-degree nodes. Indeed $(w(u,v')+w(v,u'))/2 = 2\leq w(u,v)=3$, but it does not hold that $(2w(u,v')+2w(v,u'))/2=4\leq w(u,v)=3$, so even if $u'$ and $v'$ are low-degree nodes, the degree of either $u'$ or $v'$ alone does not suffice to determine whether the contribution from an edge out of either $u$ or $v$ may be doubled.} \label{fig:highdegreecounterex} \end{figure} Figure~\ref{fig:highdegreecounterex} illustrates the base case of the proof. This also shows why it is not possible to improve the bound by low-degree considerations for the edges chosen for the lower bound. For instance, it is not true that since $v'$ is not a high-degree vertex, the contribution from edge $(u,v')$ can be counted twice. This is only possible if \emph{all} vertices adjacent to $u$ are low degree. If we can keep an estimate of the maximum degree of any adjacent vertex to $v$ for all free vertices $v\in F$ it is easy to determine whether the lower bound contribution from high-degree vertex $v$ can be multiplied by two, namely if this maximum degree is smaller than $f_1$. The estimate can be the static maximum degrees in $G$, and can be updated when a connected component contribution is computed (see next section). For each new subproblem, we can update the $T$ contribution in $O(\deg(v))$ time steps, amortized over all vertices such that the total time spent is $O(n+m)$ steps. To do this we maintain for each free vertex of $(U_0,U_1)$, a) its \emph{free degree}, b) an index of edges scanned so far, c) a count of seen (free) edges, and finally d) the total weight of the seen edges. In total, four counts are maintained per free vertex and for each subset $V_0\setminus U_0$ and $V_1\setminus U_1$, making this relatively expensive in terms of space needed per subproblem. The free degree $\deg'(v)$ of vertex $v$ in $(U_0,U_1)$ is the number of adjacent edges to free vertices (and is the degree of $v$ in the induced subgraph). Initially, the free degree of vertex $v$ is just its degree; the free degree is decreased each time a neighbor of $v$ is assigned to a subset. The count of seen edges shall be maintained as $\max(\deg'(v)-f_i+1,0)\geq 0$ for each free node $v$, and we maintain also the sum of the weights of these free edges. Note that the number of seen edges for some free node $v$ (and their weight) may have to be updated both as a result of an edge $(v,u)$ out of $v$ becoming assigned, or by some other node becoming assigned to subset $V_i$. The three remaining invariant properties of the counts are now maintained as follows. When a branching vertex $u$ is assigned to a subset, either of $f_i$ is decreased by one; thus, for each free vertex one more vertex must be counted as seen, and this is accomplished by scanning edges of the vertex until an edge whose endpoint is not fixed is met. The weight of this edge is added to the total weight of seen edges. This is eventually the lower bound contribution of the vertex. For vertices whose free degree $\deg'(v)$ decrease (that is, free vertices adjacent to the branching vertex $u$) there are two cases. If edge $(v,u)$ has already been seen, that is if edge $(v,u)$ is indexed before the currently scanned edge of $v$, then the weight of the edge can simply be subtracted from the total weight of seen edges of $v$. Since we know the position $i$ of vertex $u$ in the adjacency list of $v$, we have to subtract if $i$ is smaller than the number of scanned edges of $v$. If on the other hand the edge $(v,u)$ has not yet been seen (and scanned), the last (highest weighted) seen edge must be made unseen (and its weight subtracted from the total weight of seen edges), which is done by scanning back from the currently scanned edge until a free edge is found. In total one forwards and at most one backwards scan is made per adjacency array. For unweighted graphs, the lower bound can be computed easily. The contribution from a (high-degree) vertex is simply $\max(0,\deg'(v)-f_0+1)$. Currently, we have only implemented the general case, which is more costly (by some constant factor) than the special, unweighted case. Since this strengthening only contributes a nontrivial lower bound increase in the presence of high-degree vertices (relative to the subproblem $(U_0,U_1)$), we only compute the contribution if the maximum degree $\deg'(v)$ of any free vertex $v$ is larger than $f_1$. We can maintain an approximation of this maximum degree, namely the maximum degree from the parent subproblem, and use this to trigger the lower bound computation. Considering just the total number of free edges gives a trivial, even weaker lower bound for the term $w(V_0\setminus U_0,V_1\setminus U_1)$. Assume there are more than $f_0(f_0-1)/2+f_1(f_1-1)/2$ edges in the subgraph of $G$ induced by the free nodes $F$ ($f_0$ and and $f_1$ being the number of free nodes to be assigned to the subsets $U_0$ and $U_1$, respectively). The sum of the weights of the least weight such edges is a lower bound, since at least that number of edges must be in the cut of any partition of $F$ into subsets of sizes $f_0$ and $f_1$. If the induced subgraph has such a large number edges, at least one will be of high degree larger than $f_0-1$ and $f_1-1$, and thus the high-degree bound described above will be at least as strong, since it counts at least as many edges of at least the same weight. \subsection{The lower bound: a large unassigned component} \label{sec:largecomponent} Assume there are no high-degree vertices in the sense discussed above. We make the observation that if there is a $k$-connected component in the graph induced by $F$ of size greater than $f_0$, then at least $k$ edges will cross in any partition of the $F$ vertices into subsets of size $f_0$ and $f_1$ where here $f_0\geq f_1$. The sum of the weights of the $k$ lightest edges in such a $k$-connected component will thus be a lower bound. More generally, the (unconstrained) minimum cut of the singly connected component of size greater than $f_0$ will be a lower bound on the size constrained cut. Let $C(U_0,U_1)$ be either the weight of the $k$ smallest edges in a $k$-connected component of size greater than $f_0$ in the subgraph induced by $F$, or the value of a minimum cut in such a singly connected component. \begin{proposition} \label{prop:component} For any given subproblem $(U_0,U_1)$ it holds that \begin{eqnarray*} C(U_0,U_1) & \leq & w(V_0\setminus U_0,V_1\setminus U_1) \end{eqnarray*} for any completion $(V_0,V_1)$. \end{proposition} \begin{proof} If there is a $k$-connected component larger than $f_0$ then there will be some nodes of this component in either subset of any partition of $F$ into subsets of sizes $f_0$ and $f_1$ (assuming $f_0\geq f_1$). At least $k$ edges of the large $k$-connected component must cross between $V_0\setminus U_0$ and $V_1\setminus U_1$. The sum of the weights of the lightest such $k$ edges are therefore a lower bound on the value of any cut. Likewise is any size-unconstrained minimum cut in the subgraph induced by the large component a lower bound. \end{proof} On the other hand, if there is a high-degree vertex in the sense explained in the last section, then there is a connected component of size at least $f_0+1$. Since the high-degree bound counts the weights of particular edges (out of the high-degree vertices), this bound is at least as strong as the connected components bound. Here we settle for only a contribution of one lightest edge of a large, connected component. The component is computed by a simple breadth-first search traversal in $O(f+m)$ time steps, which can be expensive compared to the other lower bound contributions. The computation is therefore triggered by maintaining an approximate size of the largest component. Again, this approximation is just the size of the largest component from the parent subproblem, and is updated when a connected components computation is performed. During the graph traversal we also update the maximum adjacent degree for each free vertex. When a branching vertex is assigned, only the component to which this vertex belongs can be affected. It would therefore be possible to redo the connected component computation for these affected vertices, and this could perhaps be of advantage for the implementation, although not in the worst case; here, suitable dynamic connected components algorithms would have to be used. This optimization is not implemented currently, since it would require extra arrays for maintaining component numbers and sizes and storing the smallest edge weight for the components. Using DFS~\cite{Tarjan72} for the graph traversal, it would be possible to compute the 2-connected components also in linear time. Even if there is no large, $k$-connected component in the induced subgraph, there may still be a lower bound contribution. For instance, if the smallest $k$-connected component is larger than $f_1$, the smaller of the vertex sets, then \emph{some} connected component will cross between $U_0\setminus V_0$ and $U_1\setminus V_1$; thus, the sum of the weights of the $k$ least cost edges over all $k$-connected components will be a lower bound on the cut value; as above the least (unconstrained) minimum cut would also be a lower bound. More generally, if it is \emph{not} possible to pack a subset of the connected components into a subset of size $f_1$, then at least one of the components must have vertices on both sides of any cut. Again, the smallest weight edge in the induced subgraph will be a lower bound on the minimum cut value, and so will the minimum of all size-unconstrained minimum cuts. Determining this contribution implies determining that a corresponding subset sum problem does not have a solution. The subset sum problem is in itself NP-hard~\cite{GareyJohnson79}, but might be small and special enough that it could make sense to attempt a solution~\cite{Pisinger99}. We have not pursued this idea further in our current implementation. A possibly stronger bound is achieved by actually computing a minimum cut value in the graph induced by the large, singly connected component. An easily implementable algorithm~\cite{StoerWagner97} runs in $O(mn+n^2\log n)$ time steps; a better, randomized algorithm~\cite{Karger00} in $O(m\log^3 n)$ steps. We have also not yet experimented with this strengthening of the lower bound. \subsection{Maintaining an upper bound} The balancing step for the lower bound of the $w(V_0\setminus U_0,V_1)+w(U_0,V_1\setminus U_1)$ term determines an explicit assignment of the free vertices to the two subsets, that is a specific completion. We can use this completion as an upper bound on the best possible solution for the subproblem $(U_0,U_1)$. Computing the cut value of this partition would take $O(n+m)$ time, and might be too expensive to do repeatedly. However, for the case where the a large connected component bound may apply, this computation could be done almost for free. Furthermore, when branching on a vertex and creating new subproblems, it is easy to determine which vertices will change in the forced completions of the new subproblems. If only few vertices change, the cut value of the completion can be updated more cheaply similarly to what is done for instance in the Lin-Kernighan heuristic~\cite{FiducciaMattheyses82,KernighanLin70,Traff06:kpartition}. This computed upper bound is a solution candidate. Also, when upper and lower bounds meet, the lower bound is tight for the subproblem $(U_0,U_1)$, and no further branching is needed. Note, that this means that there are no edges in the cut $(V_0\setminus U_0,V_1\setminus U_1)$ for the particular completion $(V_0,V_1)$ induced by the rebalancing lower bound. We have not implemented this potential improvement so far. \subsection{Completion and branching rules} A subproblem is essentially solved if either $|U_0|=s_0$ or $|U_1|=s_1$: all free vertices can be assigned to the other subset. Furthermore, if only one vertex is missing from, e.g., $U_0$, then the vertex which has the smallest $D_1[v]$ value to the other subset plus the smallest sum of free edges (which will all cross the cut) can be assigned to $V_0$, and the remaining free vertices to subset $V_1$. No other assignment can lead to a smaller cut value of the completion $(V_0,V_1)$. Another completion rule follows from the observation that the completion implied by the lower bound with rebalancing (Proposition~\ref{prop:rebalancing}) is an optimal solution if all free vertices have free degree zero. This can easily be checked, and the corresponding solution generated; this is also implemented, and led to a reduction of a few (tens of) subproblems to be explored; since this comes at virtually no cost (it requires only maintaining the number of degree zero free vertices), this check and completion is always done when rebalancing is enabled. More generally, if it can be inferred that in this completion, there are no edges between sets $V_0\setminus U_0$ and $V_1\setminus U_1$, the completion is optimal. Other observations allows to reduce the worst-case number of subproblems that needs to be generated. We state two such observations: \begin{enumerate} \item For unassigned vertices with no free edges, not all possible assignments need to be checked. In particular, if there are $n$ such vertices, only $n+1$ of the possible $2^n$ assignments can lead to an completion value. The $i$th such subproblem for $i=0,\ldots n$ would assign the $i$ nodes with the smallest value of $\delta(v)=D_1[v_i]-D_0[v_i]$ to $V_0$, and the remaining $n-i$ nodes to $V_1$. Since these nodes have no free edges, the only contribution to the cut can come from the edges to the assigned vertices in $U_0$ or $U_1$ as counted in $D_1[v]$ and $D_0[v]$. Swapping a vertex thus assigned to $V_0$ would lead to a larger cut value. \item If there is a free edge between two nodes $u$ and $v$ each with degree one, the contribution to the cut of a completion is determined by $w(u,v)$ and the weight of the edges to the assigned vertices in $U_0$ and $U_1$. Only one of the subproblems $(U_0\cup\{u\},U_1\cup\{v\})$ or $(U_0\cup\{v\},U_1\cup\{u\})$ can lead to an optimal completion, namely the one with the smallest $D_0[u]+D_1[v]$ or $D_0[v]+D_1[u]$ value. Therefore, only 3 instead of 4 possible subproblems must be generated. \item In general, for a $k$-clique only $k+1$ instead of $2^k$ subproblems needs to be generated. \end{enumerate} These (and other, similar) observations can be used as $n$-way branching rules, instead of the binary branching rule that just generates two subproblems by assigning the chosen branching vertex to either $U_0$ or $U_1$. In our experiments, none of the first two rules above gave an advantage, and were often detrimental in that too many subproblems were generated too early. Thus, the benefit, if any, is not clear at the moment, and such branching rules have not been considered further here. \section{Solving the weighted graph bipartitioning problem} We use the task-parallel Pheet C++ framework as a general framework to implement branch-and-bound algorithms. This is described extensively in~\cite{Wimmer14:diss}, and briefly in~\cite{Traff13:strategies,Traff14:priosched}. The basic idea is to represent subproblems as tasks that can be executed in parallel when enough have been created, and let the framework take care of the selection of tasks in a priority-respecting order. To this end, Pheet supports \emph{scheduling strategies} where tasks can be spawned with an associated priority. A Pheet branch-and-bound task is shown in Figure~\ref{fig:bbtask}. When a task is processed, it is first checked that the task's lower bound is still smaller than the currently best, feasible solution (a better solution could have been found between the time the task was spawned and the time it is being processed). The computed branching vertex is used to split the subproblem into two (or more, but this is not shown here) new subproblems. For either, it is checked whether it can already be completed, and in that case whether it has lead to a new, better, global solution. If not, a new task is spawned with some computed priority. The Pheet framework will ensure that the subproblem is eventually processed by some available processing unit, preferably in good (but possibly relaxed) priority order. \begin{figure} \begin{lstlisting}[mathescape=true,columns=flexible] template <class Pheet, template <class P, class SubProblem> class Logic, template <class P, class SubProblem> class SchedulingStrategy, size_t MaxSize> void StrategyBBGraphBipartitioningTask<Pheet, Logic, SchedulingStrategy, MaxSize>:: operator()() { if(sub_problem->get_lower_bound() >= sub_problem->get_global_upper_bound()) { pc.num_irrelevant_tasks.incr(); return; } SubProblem* sub_problem2 = sub_problem->split(pc.subproblem_pc); if(sub_problem->can_complete(pc.subproblem_pc)) { sub_problem->complete_solution(pc.subproblem_pc); sub_problem->update_solution(best, pc.subproblem_pc); } else if(sub_problem->get_lower_bound() < sub_problem->get_global_upper_bound()) { Pheet::template spawn_prio<Self>(strategy(sub_problem), sub_problem, best, pc); sub_problem = NULL; } if(sub_problem2->can_complete(pc.subproblem_pc)) { sub_problem2->complete_solution(pc.subproblem_pc); sub_problem2->update_solution(best, pc.subproblem_pc); delete sub_problem2; } else if(sub_problem2->get_lower_bound() < sub_problem2->get_global_upper_bound()) { Pheet::template spawn_prio<Self>(strategy(sub_problem2), sub_problem2, best, pc); } else { delete sub_problem2; } } \end{lstlisting} \caption{A Pheet branch-and-bound task with binary branching.} \label{fig:bbtask} \end{figure} \subsection{Initial subproblems and upper bound} Branch-and-bound algorithms can benefit immensely from having a good initial feasible solution or upper bound. In our current implementation we use a simple, greedy strategy (corresponding to one iteration of the minimum cut algorithm in~\cite{StoerWagner97}) to produce an initial solution. Using a standard heuristic package like METIS~\cite{karypis11} or SCOTCH~\cite{ChevalierPellegrini08} would be a natural possibility to get a strong, initial upper bound. Easily computable solutions as provided by variations of the Lin-Kernighan heuristic are another possibility~\cite{KernighanLin70,Traff06:kpartition}. \subsection{Choosing good subproblems in parallel} Branch-and-bound normally consider subproblems in some prioritized order, with problems that are likely to lead to an improved solution or to being cut off being preferred to other problems. The subproblem priority order can have a large influence on the concrete performance of the branch-and-bound procedure, even if there is no worst-case difference. Here, we prioritize in by the difference lower bound and an estimated upper bound, such that subproblems that are close to their upper bound will be processed early. Other possibilities might be worthwhile to explore. The Pheet framework supports the possibility of prioritizing tasks and processes tasks in (relaxed) priority order. Pheet relies on various, relaxed, concurrent priority queues for this, which can provide certain semantic and performance guarantees. We refer to~\cite{Wimmer14:diss} for definitions of such semantics as well as algorithmic and implementation details. \subsection{Branching rules} For the choice of branching vertex for each subproblem there a likewise many possibilities, and the choice of branching rule can likewise have a large effect on practical performance. In our current implementation we branch on the vertex that will lead to the estimated largest increase in the lower bound when assigned to either of the subsets. \section{Benchmark results} We now present a selection of benchmark results for solving graph bipartitioning problems using the Pheet framework with the various lower bound contributions developed in the previous sections. The experiments reported here are for simple, Erd\"os-R\'enyi random graphs as in~\cite{Wimmer14:diss}. The graphs have $n$ nodes, and edges are chosen with a given, uniform probability. The bounds and the framework should be tested with standard test instances, for instance those used in~\cite{DellingGoldbergRazenshteynWerneck11,DellingGoldbergRazenshteynWerneck12,DellingWerneck12}. The graphs are either weighted, in which case edge weights are chosen uniformly at random with $w\in [1,1000]$; or unweighted, which we achieve by choosing weights $w\in [1,1]$. We give results for sparser graphs with edge probability $0.1$ (which is, asymptotically, of course rather dense), medium dense graphs with edge probability $0.5$, and dense graphs with edge probability $0.75$, and finally complete graphs with edge probability $1$. \subsection{Lower bound contributions} We first investigate the difference between the different lower bound contributions described in Section~\ref{sec:basicrebalance}, Section~\ref{sec:highdegree} and Section~\ref{sec:largecomponent}. To do this, we solve the benchmark problems sequentially, using a depth-first (non-prioritized) order on the generated subproblems. We record the time to solution and relate that to the number of subproblems that were explored. Starting from the trivial lower bound we track the reduction in number of subproblems and hopefully proportional reduction in running time by gradually strengthening the lower bound by adding the rebalancing contribution (Proposition~\ref{prop:rebalancing}), the high-degree vertex contribution (Proposition~\ref{prop:highdegree}), and the connected components contribution (Proposition~\ref{prop:component}). We give results from 5 differently generated random graphs (in Pheet with seeds $0,1,2,3,4$) from each of the four categories. In addition to the total time to solution and the number of explored subproblems, we also give the number of times a new solution was found, and the time at which the optimal (last) solution was found. For the cases where the optimal solution is found late, a better, initial solution could be of help. To check this, we also ran the experiments using the optimal solution as initial solution; this gives an objective count of how many problems must be explored to prove optimality, and is thus indicative of the strength (or weakness) of the lower bound (for the given branching rule; a different branching rule could change this; the experiment is not sensitive to the choice of subproblem priority); we only did this experiment for the strongest version of the lower bound, and here we did not measure the actual running time; we just list the, in most cases, smaller number of explored subproblems. The sequential experiments were carried out on an Intel-based desktop computer with a 4-core 3.4GHz Intel i7-2600 processor. We used \texttt{gcc 4.7.2} under Debian 4.7.2-5 Linux. All running times in seconds, and the times recorded here for a single run only; on an unloaded desktop the running times appear rather stable. The running times are only indicative; whereas the various subproblem counts are deterministic and exactly reproducible. \subsubsection{Sparse graphs} \begin{table} \begin{center} \begin{tabular}{|rrr|r|r|rr|r|} \hline $n$ & Prob. & $\max w$ & Time & Cut & Solutions & Subproblems & Opt.\ Time \\ \hline 40 & 0.1 & 1000 & 0.027267 & 7770 & 7 & 24662 & 0.026160 \\ 40 & 0.1 & 1000 & 0.065237 & 9439 & 13 & 69439 & 0.063928 \\ 40 & 0.1 & 1000 & 0.010008 & 5121 & 5 & 12520 & 0.009737 \\ 40 & 0.1 & 1000 & 0.013096 & 6523 & 10 & 16627 & 0.012922 \\ 40 & 0.1 & 1000 & 0.007524 & 6883 & 6 & 9655 & 0.005304 \\ 50 & 0.1 & 1000 & 0.320722 & 12829 & 18 & 380379 & 0.313434 \\ 50 & 0.1 & 1000 & 0.85727 & 14461 & 24 & 1012481 & 0.834905 \\ 50 & 0.1 & 1000 & 0.450444 & 9096 & 23 & 529688 & 0.433826 \\ 50 & 0.1 & 1000 & 0.343937 & 10150 & 20 & 391779 & 0.325539 \\ 50 & 0.1 & 1000 & 0.862478 & 12438 & 20 & 942418 & 0.854184 \\ 60 & 0.1 & 1000 & 12.0744 & 17502 & 29 & 11686971 & 11.278630 \\ 60 & 0.1 & 1000 & 32.5406 & 22283 & 24 & 29141096 & 30.874791 \\ 60 & 0.1 & 1000 & 6.74624 & 14585 & 20 & 6517009 & 6.500285 \\ 60 & 0.1 & 1000 & 16.389 & 14794 & 26 & 15644183 & 16.296062 \\ 60 & 0.1 & 1000 & 24.7185 & 20752 & 38 & 22683390 & 21.663690 \\ \hline 40 & 0.1 & 1 & 0.036729 & 19 & 5 & 45891 & 0.015383 \\ 40 & 0.1 & 1 & 0.044857 & 24 & 2 & 57338 & 0.041048 \\ 40 & 0.1 & 1 & 0.005785 & 14 & 1 & 8223 & 0.000006 \\ 40 & 0.1 & 1 & 0.075284 & 17 & 12 & 93339 & 0.074613 \\ 40 & 0.1 & 1 & 0.018821 & 17 & 4 & 23674 & 0.011489 \\ 50 & 0.1 & 1 & 0.73198 & 28 & 6 & 781139 & 0.711416 \\ 50 & 0.1 & 1 & 2.83857 & 32 & 10 & 3056976 & 2.755318 \\ 50 & 0.1 & 1 & 0.515393 & 25 & 2 & 626003 & 0.419959 \\ 50 & 0.1 & 1 & 1.20943 & 25 & 11 & 1273991 & 1.096330 \\ 50 & 0.1 & 1 & 4.80033 & 33 & 14 & 4904437 & 4.631338 \\ 60 & 0.1 & 1 & 21.8007 & 42 & 6 & 19637124 & 16.885930 \\ 60 & 0.1 & 1 & 97.601 & 53 & 10 & 84588305 & 75.646062 \\ 60 & 0.1 & 1 & 32.8037 & 39 & 12 & 29165329 & 22.103448 \\ 60 & 0.1 & 1 & 9.48105 & 35 & 5 & 8931547 & 9.305966 \\ 60 & 0.1 & 1 & 126.377 & 50 & 11 & 113479717 & 112.269199 \\ \hline \end{tabular} \end{center} \caption{Sparse random graphs with edge probability $0.1$, $w\in[1,1000]$ and $w\in[1,1]$. Trivial lower bound.} \label{tab:sparse-norebal} \end{table} \begin{table} \begin{center} \begin{tabular}{|rrr|r|r|rr|r|} \hline $n$ & Prob. & $\max w$ & Time & Cut & Solutions & Subproblems & Opt.\ Time \\ \hline 40 & 0.1 & 1000 & 0.007509 & 7770 & 7 & 3164 & 0.005886 \\ 40 & 0.1 & 1000 & 0.006289 & 9439 & 13 & 2888 & 0.004836 \\ 40 & 0.1 & 1000 & 0.001719 & 5121 & 5 & 735 & 0.001296 \\ 40 & 0.1 & 1000 & 0.003596 & 6523 & 10 & 1843 & 0.003427 \\ 40 & 0.1 & 1000 & 0.002714 & 6883 & 6 & 1178 & 0.000321 \\ 50 & 0.1 & 1000 & 0.057015 & 12829 & 18 & 26314 & 0.052750 \\ 50 & 0.1 & 1000 & 0.078241 & 14461 & 24 & 45226 & 0.071687 \\ 50 & 0.1 & 1000 & 0.03234 & 9096 & 23 & 20536 & 0.021565 \\ 50 & 0.1 & 1000 & 0.031847 & 10150 & 20 & 19785 & 0.017051 \\ 50 & 0.1 & 1000 & 0.039503 & 12438 & 20 & 23618 & 0.038913 \\ 60 & 0.1 & 1000 & 0.160308 & 17502 & 29 & 77425 & 0.053616 \\ 60 & 0.1 & 1000 & 0.771585 & 22283 & 24 & 394574 & 0.421817 \\ 60 & 0.1 & 1000 & 0.160184 & 14585 & 20 & 77410 & 0.065352 \\ 60 & 0.1 & 1000 & 0.327008 & 14794 & 26 & 164106 & 0.273994 \\ 60 & 0.1 & 1000 & 1.23954 & 20752 & 38 & 641762 & 0.544033 \\ \hline 40 & 0.1 & 1 & 0.004571 & 19 & 5 & 3678 & 0.000330 \\ 40 & 0.1 & 1 & 0.004976 & 24 & 2 & 4116 & 0.003736 \\ 40 & 0.1 & 1 & 0.001199 & 14 & 1 & 899 & 0.000010 \\ 40 & 0.1 & 1 & 0.004194 & 17 & 12 & 3482 & 0.003519 \\ 40 & 0.1 & 1 & 0.002543 & 17 & 4 & 1965 & 0.000275 \\ 50 & 0.1 & 1 & 0.04406 & 28 & 6 & 30093 & 0.041470 \\ 50 & 0.1 & 1 & 0.096952 & 32 & 10 & 67991 & 0.074959 \\ 50 & 0.1 & 1 & 0.088082 & 25 & 2 & 60791 & 0.044789 \\ 50 & 0.1 & 1 & 0.037535 & 25 & 11 & 25168 & 0.008754 \\ 50 & 0.1 & 1 & 0.103074 & 33 & 14 & 70069 & 0.068494 \\ 60 & 0.1 & 1 & 0.236067 & 42 & 6 & 128962 & 0.029440 \\ 60 & 0.1 & 1 & 3.96175 & 53 & 10 & 2410554 & 1.395450 \\ 60 & 0.1 & 1 & 0.990479 & 39 & 12 & 558587 & 0.075503 \\ 60 & 0.1 & 1 & 0.251598 & 35 & 5 & 139853 & 0.209421 \\ 60 & 0.1 & 1 & 4.33112 & 50 & 11 & 2599774 & 2.396716 \\ \hline \end{tabular} \end{center} \caption{Sparse random graphs with edge probability $0.1$, $w\in[1,1000]$ and $w\in[1,1]$. Lower bound with rebalancing contribution.} \label{tab:sparse-rebal} \end{table} \begin{table} \begin{center} \begin{tabular}{|rrr|r|r|rr|r|} \hline $n$ & Prob. & $\max w$ & Time & Cut & Solutions & Subproblems & Opt.\ Time \\ \hline 40 & 0.1 & 1000 & 0.004721 & 7770 & 7 & 3154 & 0.003703 \\ 40 & 0.1 & 1000 & 0.004272 & 9439 & 13 & 2881 & 0.003260 \\ 40 & 0.1 & 1000 & 0.001172 & 5121 & 5 & 724 & 0.000881 \\ 40 & 0.1 & 1000 & 0.00273 & 6523 & 10 & 1837 & 0.002567 \\ 40 & 0.1 & 1000 & 0.001854 & 6883 & 6 & 1178 & 0.000302 \\ 50 & 0.1 & 1000 & 0.048337 & 12829 & 17 & 26293 & 0.044498 \\ 50 & 0.1 & 1000 & 0.080309 & 14461 & 24 & 45197 & 0.073267 \\ 50 & 0.1 & 1000 & 0.035161 & 9096 & 23 & 20489 & 0.023535 \\ 50 & 0.1 & 1000 & 0.03498 & 10150 & 19 & 19772 & 0.018809 \\ 50 & 0.1 & 1000 & 0.043134 & 12438 & 20 & 23592 & 0.042481 \\ 60 & 0.1 & 1000 & 0.17328 & 17502 & 29 & 77404 & 0.058439 \\ 60 & 0.1 & 1000 & 0.842746 & 22283 & 24 & 394561 & 0.458983 \\ 60 & 0.1 & 1000 & 0.174728 & 14585 & 20 & 77410 & 0.071872 \\ 60 & 0.1 & 1000 & 0.3576 & 14794 & 23 & 163998 & 0.299697 \\ 60 & 0.1 & 1000 & 1.35194 & 20752 & 38 & 641762 & 0.594848 \\ \hline 40 & 0.1 & 1 & 0.005176 & 19 & 5 & 3673 & 0.000348 \\ 40 & 0.1 & 1 & 0.005727 & 24 & 2 & 4113 & 0.004282 \\ 40 & 0.1 & 1 & 0.00137 & 14 & 1 & 899 & 0.000006 \\ 40 & 0.1 & 1 & 0.004799 & 17 & 12 & 3453 & 0.004016 \\ 40 & 0.1 & 1 & 0.002827 & 17 & 4 & 1965 & 0.000286 \\ 50 & 0.1 & 1 & 0.049363 & 28 & 6 & 30063 & 0.046491 \\ 50 & 0.1 & 1 & 0.109406 & 32 & 10 & 67970 & 0.084869 \\ 50 & 0.1 & 1 & 0.098004 & 25 & 2 & 60791 & 0.049769 \\ 50 & 0.1 & 1 & 0.041763 & 25 & 11 & 25163 & 0.010042 \\ 50 & 0.1 & 1 & 0.113649 & 33 & 14 & 70034 & 0.075498 \\ 60 & 0.1 & 1 & 0.261038 & 42 & 6 & 128950 & 0.032600 \\ 60 & 0.1 & 1 & 4.3952 & 53 & 10 & 2410551 & 1.544418 \\ 60 & 0.1 & 1 & 1.08553 & 39 & 12 & 558587 & 0.084374 \\ 60 & 0.1 & 1 & 0.280103 & 35 & 5 & 139844 & 0.233175 \\ 60 & 0.1 & 1 & 4.77637 & 50 & 11 & 2599761 & 2.651952 \\ \hline \end{tabular} \end{center} \caption{Sparse random graphs with edge probability $0.1$, $w\in[1,1000]$ and $w\in[1,1]$. Lower bound with rebalancing and high-degree contributions.} \label{tab:sparse-highdegree} \end{table} \begin{table} \begin{center} \begin{tabular}{|rrr|r|r|rrr|r|} \hline $n$ & Prob. & $\max w$ & Time & Cut & Solutions & Subproblems & With optimal & Opt.\ Time \\ \hline 40 & 0.1 & 1000 & 0.007153 & 7770 & 7 & 3121 & 1701 & 0.005560 \\ 40 & 0.1 & 1000 & 0.006443 & 9439 & 13 & 2842 & 1228 & 0.004901 \\ 40 & 0.1 & 1000 & 0.001643 & 5121 & 5 & 719 & 280 & 0.001208 \\ 40 & 0.1 & 1000 & 0.003775 & 6523 & 10 & 1837 & 1260 & 0.003540 \\ 40 & 0.1 & 1000 & 0.002586 & 6883 & 6 & 1144 & 1051 & 0.000390 \\ 50 & 0.1 & 1000 & 0.077288 & 12829 & 17 & 25850 & 13812 & 0.070880 \\ 50 & 0.1 & 1000 & 0.1288 & 14461 & 24 & 44522 & 24324 & 0.117323 \\ 50 & 0.1 & 1000 & 0.051628 & 9096 & 23 & 20021 & 10235 & 0.033532 \\ 50 & 0.1 & 1000 & 0.054827 & 10150 & 19 & 19490 & 11494 & 0.028375 \\ 50 & 0.1 & 1000 & 0.069346 & 12438 & 20 & 23157 & 7214 & 0.068327 \\ 60 & 0.1 & 1000 & 0.288285 & 17502 & 29 & 75860 & 51678 & 0.091989 \\ 60 & 0.1 & 1000 & 1.47682 & 22283 & 24 & 392842 & 250649 & 0.796601 \\ 60 & 0.1 & 1000 & 0.285672 & 14585 & 20 & 76669 & 47630 & 0.115380 \\ 60 & 0.1 & 1000 & 0.600668 & 14794 & 23 & 162745 & 66431 & 0.501793 \\ 60 & 0.1 & 1000 & 2.26215 & 20752 & 38 & 627491 & 426709 & 0.984418 \\ \hline 40 & 0.1 & 1 & 0.005906 & 19 & 5 & 2907 & 2669 & 0.000434 \\ 40 & 0.1 & 1 & 0.006806 & 24 & 2 & 3242 & 2218 & 0.005147 \\ 40 & 0.1 & 1 & 0.001549 & 14 & 1 & 866 & 866 & 0.000008 \\ 40 & 0.1 & 1 & 0.005656 & 17 & 12 & 3195 & 981 & 0.004748 \\ 40 & 0.1 & 1 & 0.00333 & 17 & 4 & 1654 & 1524 & 0.000328 \\ 50 & 0.1 & 1 & 0.059948 & 28 & 6 & 22192 & 13202 & 0.056434 \\ 50 & 0.1 & 1 & 0.134392 & 32 & 10 & 49507 & 15853 & 0.103304 \\ 50 & 0.1 & 1 & 0.114136 & 25 & 2 & 47341 & 37054 & 0.057939 \\ 50 & 0.1 & 1 & 0.050109 & 25 & 11 & 18543 & 13614 & 0.011498 \\ 50 & 0.1 & 1 & 0.135859 & 33 & 14 & 50585 & 24836 & 0.089800 \\ 60 & 0.1 & 1 & 0.348262 & 42 & 6 & 96073 & 84417 & 0.040723 \\ 60 & 0.1 & 1 & 5.77289 & 53 & 10 & 1728577 & 1526818 & 2.003938 \\ 60 & 0.1 & 1 & 1.31056 & 39 & 12 & 404049 & 366201 & 0.095397 \\ 60 & 0.1 & 1 & 0.354385 & 35 & 5 & 100256 & 42995 & 0.294146 \\ 60 & 0.1 & 1 & 6.1523 & 50 & 11 & 1873524 & 1172829 & 3.360075 \\ \hline \end{tabular} \end{center} \caption{Sparse random graphs with edge probability $0.1$, $w\in[1,1000]$ and $w\in[1,1]$. Lower bound with rebalancing, high-degree and large connected component contributions.} \label{tab:sparse-all} \end{table} The results for sparse graphs are given in Table~\ref{tab:sparse-norebal}, Table~\ref{tab:sparse-rebal}, Table~\ref{tab:sparse-highdegree} and Table~\ref{tab:sparse-all}. We first notice (and this observation holds also for the other graph categories) that the rebalancing lower bound contribution leads to a huge reduction in number of subproblems; for the sparse graphs often more than a factor of 20, and both for weighted and unweighted problems. A similar reduction in running times follows. The high-degree bound has, as would be expected, no effect here, the number of subproblems is for all graphs the same. Fortunately, running times seem to increase only slightly, which could mean that the larger memory space needed to represent the subproblems for this bound contribution is not in itself too costly. Adding the large connected components contribution reduces the number of subproblems that have to be considered by a significant factor less than 2, especially for the unweighted graphs (as could be hoped for). Unfortunately, the extra cost for repeatedly computing connected components outweigh the reduction in number of subproblems, resulting in an increase in time to solution by a small factor less than 2. As can be seen in Table~\ref{tab:sparse-all}, the optimal solution is often found late, about half-way through, and knowing the optimal solution as expected leads to a significant reduction in numbers of subproblems that must be explored; the reduction is less than a factor of 2, though. \subsubsection{Medium dense graphs} \begin{table} \begin{center} \begin{tabular}{|rrr|r|r|rr|r|} \hline $n$ & Prob. & $\max w$ & Time & Cut & Solutions & Subproblems & Opt.\ Time \\ \hline 35 & 0.5 & 1000 & 1.52957 & 58764 & 8 & 1631423 & 1.442220 \\ 35 & 0.5 & 1000 & 1.45441 & 53759 & 13 & 1569317 & 1.090717 \\ 35 & 0.5 & 1000 & 2.07381 & 57403 & 5 & 2281700 & 0.949366 \\ 35 & 0.5 & 1000 & 1.00532 & 52375 & 12 & 1072568 & 0.479008 \\ 35 & 0.5 & 1000 & 1.35941 & 51263 & 6 & 1542516 & 0.805244 \\ 40 & 0.5 & 1000 & 8.13641 & 77452 & 7 & 7398960 & 7.119544 \\ 40 & 0.5 & 1000 & 4.19626 & 65643 & 10 & 3916431 & 2.272946 \\ 40 & 0.5 & 1000 & 7.92397 & 74034 & 17 & 7399920 & 6.992345 \\ 40 & 0.5 & 1000 & 9.02878 & 69479 & 20 & 8579493 & 6.311949 \\ 40 & 0.5 & 1000 & 3.27055 & 64952 & 8 & 2982317 & 2.088845 \\ 45 & 0.5 & 1000 & 212.693 & 97409 & 13 & 173104481 & 109.589662 \\ 45 & 0.5 & 1000 & 207.349 & 88328 & 15 & 182233817 & 171.094763 \\ 45 & 0.5 & 1000 & 289.602 & 99578 & 9 & 244147116 & 254.704806 \\ 45 & 0.5 & 1000 & 159.806 & 87290 & 24 & 136813957 & 115.479522 \\ 45 & 0.5 & 1000 & 98.8471 & 82358 & 12 & 84298381 & 54.957056 \\ 50 & 0.5 & 1000 & 1677.05 & 122708 & 13 & 1244044666 & 1270.327979 \\ 50 & 0.5 & 1000 & 757.332 & 113679 & 13 & 568874993 & 31.799608 \\ 50 & 0.5 & 1000 & 1654.28 & 119443 & 26 & 1190261679 & 1008.096581 \\ 50 & 0.5 & 1000 & 1034.73 & 110783 & 10 & 785507028 & 886.124273 \\ 50 & 0.5 & 1000 & 666.77 & 106336 & 14 & 509060437 & 446.841404 \\ \hline 35 & 0.5 & 1 & 4.65021 & 124 & 6 & 5269158 & 3.204704 \\ 35 & 0.5 & 1 & 4.21001 & 120 & 7 & 4742440 & 1.624278 \\ 35 & 0.5 & 1 & 4.2815 & 122 & 4 & 4667670 & 0.311393 \\ 35 & 0.5 & 1 & 3.33108 & 118 & 4 & 3663708 & 1.444637 \\ 35 & 0.5 & 1 & 4.15371 & 116 & 7 & 4665842 & 1.641493 \\ 40 & 0.5 & 1 & 30.91 & 164 & 4 & 29116782 & 16.022385 \\ 40 & 0.5 & 1 & 21.1244 & 152 & 2 & 20240985 & 19.215322 \\ 40 & 0.5 & 1 & 25.9468 & 164 & 5 & 24385212 & 14.307037 \\ 40 & 0.5 & 1 & 25.9925 & 155 & 5 & 24769899 & 10.099783 \\ 40 & 0.5 & 1 & 19.4168 & 148 & 6 & 18709834 & 9.208803 \\ 45 & 0.5 & 1 & 848.796 & 208 & 11 & 727360976 & 364.433880 \\ 45 & 0.5 & 1 & 567.632 & 194 & 5 & 512035561 & 187.807958 \\ 45 & 0.5 & 1 & 840.502 & 212 & 6 & 729022051 & 556.588036 \\ 45 & 0.5 & 1 & 475.761 & 196 & 3 & 413970887 & 62.586686 \\ 45 & 0.5 & 1 & 539.03 & 189 & 10 & 475797670 & 122.222540 \\ 50 & 0.5 & 1 & 6641.34 & 259 & 4 & 826628810 & 2926.981140 \\ 50 & 0.5 & 1 & 7660.52 & 253 & 11 & 1771570933 & 5053.982181 \\ 50 & 0.5 & 1 & 6319.32 & 263 & 8 & 471230838 & 2461.749443 \\ 50 & 0.5 & 1 & 5152.95 & 244 & 9 & -202499879 & 3026.881798 \\ 50 & 0.5 & 1 & 3591.88 & 234 & 14 & -1543394689 & 2380.251020 \\ \hline \end{tabular} \end{center} \caption{Medium random graphs with edge probability $0.5$, $w\in[1,1000]$ and $w\in[1,1]$. Trivial lower bound.} \label{tab:medium-norebal} \end{table} \begin{table} \begin{center} \begin{tabular}{|rrr|r|r|rr|r|} \hline $n$ & Prob. & $\max w$ & Time & Cut & Solutions & Subproblems & Opt.\ Time \\ \hline 35 & 0.5 & 1000 & 0.264013 & 58764 & 8 & 240742 & 0.260441 \\ 35 & 0.5 & 1000 & 0.248807 & 53759 & 13 & 247377 & 0.203659 \\ 35 & 0.5 & 1000 & 0.376443 & 57403 & 5 & 385777 & 0.184042 \\ 35 & 0.5 & 1000 & 0.168552 & 52375 & 12 & 157670 & 0.077293 \\ 35 & 0.5 & 1000 & 0.273458 & 51263 & 6 & 270768 & 0.173758 \\ 40 & 0.5 & 1000 & 0.954789 & 77452 & 7 & 804652 & 0.830155 \\ 40 & 0.5 & 1000 & 0.499157 & 65643 & 10 & 407117 & 0.194435 \\ 40 & 0.5 & 1000 & 0.863879 & 74034 & 17 & 745162 & 0.729009 \\ 40 & 0.5 & 1000 & 1.32082 & 69479 & 20 & 1124210 & 0.848951 \\ 40 & 0.5 & 1000 & 0.210091 & 64952 & 8 & 164562 & 0.091707 \\ 45 & 0.5 & 1000 & 16.2536 & 97409 & 13 & 12920447 & 9.249073 \\ 45 & 0.5 & 1000 & 20.8354 & 88328 & 15 & 17088208 & 18.617669 \\ 45 & 0.5 & 1000 & 21.8805 & 99578 & 9 & 17823261 & 20.776830 \\ 45 & 0.5 & 1000 & 14.1073 & 87290 & 24 & 11424168 & 11.470726 \\ 45 & 0.5 & 1000 & 6.7862 & 82358 & 12 & 5273280 & 4.230401 \\ 50 & 0.5 & 1000 & 92.208 & 122708 & 13 & 69114119 & 64.386391 \\ 50 & 0.5 & 1000 & 39.5426 & 113679 & 13 & 28419256 & 0.508509 \\ 50 & 0.5 & 1000 & 83.9424 & 119443 & 26 & 63410193 & 42.537509 \\ 50 & 0.5 & 1000 & 60.9577 & 110783 & 10 & 44579202 & 49.544491 \\ 50 & 0.5 & 1000 & 40.4662 & 106336 & 14 & 29260832 & 24.309945 \\ \hline 35 & 0.5 & 1 & 0.84897 & 124 & 6 & 941536 & 0.646175 \\ 35 & 0.5 & 1 & 0.691073 & 120 & 7 & 749916 & 0.279613 \\ 35 & 0.5 & 1 & 0.737433 & 122 & 4 & 799902 & 0.047950 \\ 35 & 0.5 & 1 & 0.473472 & 118 & 4 & 512213 & 0.225050 \\ 35 & 0.5 & 1 & 0.72023 & 116 & 7 & 804766 & 0.309161 \\ 40 & 0.5 & 1 & 3.13725 & 164 & 4 & 3007647 & 1.411749 \\ 40 & 0.5 & 1 & 2.19887 & 152 & 2 & 2049958 & 2.011402 \\ 40 & 0.5 & 1 & 2.6435 & 164 & 5 & 2485163 & 1.317400 \\ 40 & 0.5 & 1 & 3.08175 & 155 & 5 & 2912912 & 0.900051 \\ 40 & 0.5 & 1 & 1.80619 & 148 & 6 & 1689634 & 0.698424 \\ 45 & 0.5 & 1 & 68.1693 & 208 & 11 & 59670375 & 31.835521 \\ 45 & 0.5 & 1 & 49.584 & 194 & 5 & 42974470 & 17.051052 \\ 45 & 0.5 & 1 & 58.1044 & 212 & 6 & 50732290 & 42.331557 \\ 45 & 0.5 & 1 & 33.5946 & 196 & 3 & 28382980 & 3.404407 \\ 45 & 0.5 & 1 & 44.7034 & 189 & 10 & 38908212 & 9.043913 \\ 50 & 0.5 & 1 & 418.594 & 259 & 4 & 344992577 & 158.152543 \\ 50 & 0.5 & 1 & 518.243 & 253 & 11 & 431767620 & 317.159786 \\ 50 & 0.5 & 1 & 347.785 & 263 & 8 & 284400088 & 120.231741 \\ 50 & 0.5 & 1 & 271.259 & 244 & 9 & 215621591 & 150.366613 \\ 50 & 0.5 & 1 & 189.26 & 234 & 14 & 148813817 & 122.718955 \\ \hline \end{tabular} \end{center} \caption{Medium random graphs with edge probability $0.5$, $w\in[1,1000]$ and $w\in[1,1]$. Lower bound with rebalancing contribution.} \label{tab:medium-rebal} \end{table} \begin{table} \begin{center} \begin{tabular}{|rrr|r|r|rr|r|} \hline $n$ & Prob. & $\max w$ & Time & Cut & Solutions & Subproblems & Opt.\ Time \\ \hline 35 & 0.5 & 1000 & 0.437692 & 58764 & 8 & 235292 & 0.430839 \\ 35 & 0.5 & 1000 & 0.425497 & 53759 & 13 & 242591 & 0.347885 \\ 35 & 0.5 & 1000 & 0.666342 & 57403 & 5 & 373692 & 0.325182 \\ 35 & 0.5 & 1000 & 0.288164 & 52375 & 12 & 154324 & 0.131943 \\ 35 & 0.5 & 1000 & 0.47135 & 51263 & 6 & 265518 & 0.301217 \\ 40 & 0.5 & 1000 & 1.67566 & 77452 & 7 & 774893 & 1.457920 \\ 40 & 0.5 & 1000 & 0.846506 & 65643 & 10 & 399509 & 0.331395 \\ 40 & 0.5 & 1000 & 1.5265 & 74034 & 17 & 728341 & 1.290223 \\ 40 & 0.5 & 1000 & 2.2194 & 69479 & 20 & 1100933 & 1.449646 \\ 40 & 0.5 & 1000 & 0.323746 & 64952 & 8 & 163934 & 0.142388 \\ 45 & 0.5 & 1000 & 29.0639 & 97409 & 13 & 12614199 & 16.531646 \\ 45 & 0.5 & 1000 & 35.5419 & 88328 & 15 & 16843898 & 31.671609 \\ 45 & 0.5 & 1000 & 38.5121 & 99578 & 9 & 17408144 & 36.496099 \\ 45 & 0.5 & 1000 & 24.9732 & 87290 & 24 & 11281722 & 20.296342 \\ 45 & 0.5 & 1000 & 11.2825 & 82358 & 12 & 5241779 & 7.054381 \\ 50 & 0.5 & 1000 & 167.633 & 122708 & 13 & 66634939 & 117.179447 \\ 50 & 0.5 & 1000 & 69.967 & 113679 & 13 & 27799777 & 0.975712 \\ 50 & 0.5 & 1000 & 159.028 & 119443 & 26 & 60947964 & 80.593920 \\ 50 & 0.5 & 1000 & 105.542 & 110783 & 10 & 43918527 & 86.097362 \\ 50 & 0.5 & 1000 & 69.2837 & 106336 & 14 & 28457488 & 42.036457 \\ \hline 35 & 0.5 & 1 & 1.35922 & 124 & 6 & 871277 & 1.033276 \\ 35 & 0.5 & 1 & 1.07911 & 120 & 7 & 716940 & 0.434534 \\ 35 & 0.5 & 1 & 1.17949 & 122 & 4 & 725398 & 0.071981 \\ 35 & 0.5 & 1 & 0.752047 & 118 & 4 & 493487 & 0.356672 \\ 35 & 0.5 & 1 & 1.16846 & 116 & 7 & 770453 & 0.503175 \\ 40 & 0.5 & 1 & 5.07885 & 164 & 4 & 2864374 & 2.286510 \\ 40 & 0.5 & 1 & 3.47258 & 152 & 2 & 2018241 & 3.173854 \\ 40 & 0.5 & 1 & 4.23372 & 164 & 5 & 2423006 & 2.115559 \\ 40 & 0.5 & 1 & 4.81996 & 155 & 5 & 2795728 & 1.407476 \\ 40 & 0.5 & 1 & 2.75878 & 148 & 6 & 1656079 & 1.090572 \\ 45 & 0.5 & 1 & 109.491 & 208 & 11 & 56006759 & 50.681042 \\ 45 & 0.5 & 1 & 78.2504 & 194 & 5 & 41889646 & 26.893112 \\ 45 & 0.5 & 1 & 95.5201 & 212 & 6 & 48662675 & 69.489188 \\ 45 & 0.5 & 1 & 52.6926 & 196 & 3 & 27582111 & 5.511826 \\ 45 & 0.5 & 1 & 68.3469 & 189 & 10 & 37849057 & 13.994667 \\ 50 & 0.5 & 1 & 679.514 & 259 & 4 & 319026234 & 254.246279 \\ 50 & 0.5 & 1 & 828.198 & 253 & 11 & 415976171 & 507.750523 \\ 50 & 0.5 & 1 & 564.802 & 263 & 8 & 256492804 & 192.467646 \\ 50 & 0.5 & 1 & 411.546 & 244 & 9 & 211109755 & 227.832552 \\ 50 & 0.5 & 1 & 293.403 & 234 & 14 & 142303065 & 189.947916 \\ \hline \end{tabular} \end{center} \caption{Medium random graphs with edge probability $0.5$, $w\in[1,1000]$ and $w\in[1,1]$. Lower bound with rebalancing and high-degree contributions.} \label{tab:medium-highdegree} \end{table} \begin{table} \begin{center} \begin{tabular}{|rrr|r|r|rrr|r|} \hline $n$ & Prob. & $\max w$ & Time & Cut & Solutions & Subproblems & With optimal & Opt.\ Time \\ \hline 35 & 0.5 & 1000 & 0.457685 & 58764 & 8 & 235288 & 168998 & 0.450484 \\ 35 & 0.5 & 1000 & 0.456458 & 53759 & 13 & 242548 & 186531 & 0.372767 \\ 35 & 0.5 & 1000 & 0.701454 & 57403 & 5 & 373655 & 357593 & 0.347343 \\ 35 & 0.5 & 1000 & 0.303973 & 52375 & 12 & 154306 & 107663 & 0.139753 \\ 35 & 0.5 & 1000 & 0.490333 & 51263 & 6 & 265476 & 224962 & 0.315305 \\ 40 & 0.5 & 1000 & 1.76439 & 77452 & 7 & 774877 & 602759 & 1.540306 \\ 40 & 0.5 & 1000 & 0.856906 & 65643 & 10 & 399493 & 381139 & 0.338882 \\ 40 & 0.5 & 1000 & 1.52563 & 74034 & 17 & 728331 & 660421 & 1.286866 \\ 40 & 0.5 & 1000 & 2.32557 & 69479 & 20 & 1100861 & 750247 & 1.510679 \\ 40 & 0.5 & 1000 & 0.366715 & 64952 & 8 & 163902 & 120719 & 0.161070 \\ 45 & 0.5 & 1000 & 30.03 & 97409 & 13 & 12614160 & 9331432 & 17.022816 \\ 45 & 0.5 & 1000 & 36.9243 & 88328 & 15 & 16843385 & 13513268 & 33.014433 \\ 45 & 0.5 & 1000 & 40.0554 & 99578 & 9 & 17408058 & 15693005 & 38.014035 \\ 45 & 0.5 & 1000 & 25.6186 & 87290 & 24 & 11281384 & 6465924 & 20.866948 \\ 45 & 0.5 & 1000 & 12.2618 & 82358 & 12 & 5240191 & 3495765 & 7.735454 \\ 50 & 0.5 & 1000 & 171.224 & 122708 & 13 & 66634689 & 55686051 & 119.794947 \\ 50 & 0.5 & 1000 & 72.0477 & 113679 & 13 & 27799722 & 27585136 & 1.010014 \\ 50 & 0.5 & 1000 & 160.152 & 119443 & 26 & 60947916 & 43152850 & 80.989391 \\ 50 & 0.5 & 1000 & 106.719 & 110783 & 10 & 43917268 & 31788303 & 87.113578 \\ 50 & 0.5 & 1000 & 70.4171 & 106336 & 14 & 28456665 & 22696796 & 42.707401 \\ \hline 35 & 0.5 & 1 & 1.4499 & 124 & 6 & 855438 & 601343 & 1.101491 \\ 35 & 0.5 & 1 & 1.20145 & 120 & 7 & 696530 & 618260 & 0.486955 \\ 35 & 0.5 & 1 & 1.23778 & 122 & 4 & 716422 & 696641 & 0.072427 \\ 35 & 0.5 & 1 & 0.854104 & 118 & 4 & 479529 & 393921 & 0.407605 \\ 35 & 0.5 & 1 & 1.29394 & 116 & 7 & 747361 & 645569 & 0.551096 \\ 40 & 0.5 & 1 & 5.50524 & 164 & 4 & 2818789 & 2344215 & 2.458640 \\ 40 & 0.5 & 1 & 4.05059 & 152 & 2 & 1959144 & 1691452 & 3.707572 \\ 40 & 0.5 & 1 & 4.71051 & 164 & 5 & 2373464 & 2187064 & 2.329712 \\ 40 & 0.5 & 1 & 5.4678 & 155 & 5 & 2720383 & 2432052 & 1.599556 \\ 40 & 0.5 & 1 & 3.40263 & 148 & 6 & 1586829 & 1236544 & 1.291652 \\ 45 & 0.5 & 1 & 118.224 & 208 & 11 & 55380489 & 46060711 & 55.248718 \\ 45 & 0.5 & 1 & 91.6279 & 194 & 5 & 40993549 & 36002769 & 30.710736 \\ 45 & 0.5 & 1 & 102.48 & 212 & 6 & 48161062 & 40186440 & 74.656017 \\ 45 & 0.5 & 1 & 62.6245 & 196 & 3 & 26867057 & 25884153 & 6.233091 \\ 45 & 0.5 & 1 & 87.8081 & 189 & 10 & 36532129 & 33329439 & 17.232088 \\ 50 & 0.5 & 1 & 737.273 & 259 & 4 & 315805991 & 291043130 & 274.235916 \\ 50 & 0.5 & 1 & 940.956 & 253 & 11 & 408914142 & 309539236 & 575.418819 \\ 50 & 0.5 & 1 & 583.4 & 263 & 8 & 255530043 & 214170641 & 198.905050 \\ 50 & 0.5 & 1 & 521.97 & 244 & 9 & 204611587 & 171705443 & 285.835977 \\ 50 & 0.5 & 1 & 362.011 & 234 & 14 & 138985134 & 98499266 & 229.681082 \\ \hline \end{tabular} \end{center} \caption{Medium random graphs with edge probability $0.5$, $w\in[1,1000]$ and $w\in[1,1]$. Lower bound with rebalancing, high-degree and large component contributions.} \label{tab:medium-all} \end{table} The results for medium dense graphs are listed in Table~\ref{tab:medium-norebal}, Table~\ref{tab:medium-rebal}, Table~\ref{tab:medium-highdegree} and Table~\ref{tab:medium-all}. Again, the benefits from the rebalancing contribution are enormous, both for weighted and unweighted case, and obviously pay off proportionally in running time (factors of 15 and more). Here, the high-degree contribution is triggered and leads to a small reduction in numbers of subproblems, but the computation is expensive and has a negative effect on the time to solution which can almost double. The same holds for the large connected components contribution, which although the number of subproblems can be reduced slightly, increases the running times by a small factor less than 2. In most cases the optimal solution is found relatively late, and there would therefore be a benefit (in number of subproblems to explore) of having a better initial solution; the effect is less than a factor of 2, though. \subsubsection{Dense graphs} \begin{table} \begin{center} \begin{tabular}{|rrr|r|r|rr|r|} \hline $n$ & Prob. & $\max w$ & Time & Cut & Solutions & Subproblems & Opt.\ Time \\ \hline 30 & 0.75 & 1000 & 0.370472 & 70916 & 11 & 360252 & 0.259478 \\ 30 & 0.75 & 1000 & 0.337704 & 68761 & 2 & 349245 & 0.083185 \\ 30 & 0.75 & 1000 & 0.221823 & 67895 & 2 & 214933 & 0.132072 \\ 30 & 0.75 & 1000 & 0.308535 & 66801 & 7 & 313060 & 0.253589 \\ 30 & 0.75 & 1000 & 0.309682 & 65323 & 11 & 317570 & 0.217466 \\ 35 & 0.75 & 1000 & 13.593 & 101149 & 7 & 12514058 & 6.705426 \\ 35 & 0.75 & 1000 & 6.52348 & 88464 & 8 & 5895137 & 4.358233 \\ 35 & 0.75 & 1000 & 5.20966 & 91901 & 7 & 4572475 & 4.766952 \\ 35 & 0.75 & 1000 & 4.94418 & 85501 & 11 & 4295028 & 3.135605 \\ 35 & 0.75 & 1000 & 6.67847 & 87457 & 22 & 5989707 & 2.272913 \\ 40 & 0.75 & 1000 & 93.225 & 131755 & 8 & 72199908 & 58.842574 \\ 40 & 0.75 & 1000 & 65.6515 & 121963 & 10 & 51272133 & 14.000955 \\ 40 & 0.75 & 1000 & 70.0444 & 126009 & 14 & 53027510 & 52.134243 \\ 40 & 0.75 & 1000 & 53.4483 & 119664 & 9 & 41483027 & 44.666847 \\ 40 & 0.75 & 1000 & 49.6401 & 114953 & 16 & 39378102 & 20.038079 \\ \hline 30 & 0.75 & 1 & 0.891637 & 147 & 3 & 979764 & 0.725699 \\ 30 & 0.75 & 1 & 1.02601 & 150 & 4 & 1128185 & 0.109506 \\ 30 & 0.75 & 1 & 0.836867 & 148 & 2 & 906552 & 0.797915 \\ 30 & 0.75 & 1 & 0.938687 & 148 & 1 & 1011748 & 0.000007 \\ 30 & 0.75 & 1 & 0.932776 & 145 & 5 & 1043643 & 0.221146 \\ 35 & 0.75 & 1 & 32.3945 & 205 & 1 & 31235221 & 0.000009 \\ 35 & 0.75 & 1 & 22.7523 & 195 & 4 & 21695745 & 12.457475 \\ 35 & 0.75 & 1 & 37.0252 & 206 & 6 & 36168861 & 24.928672 \\ 35 & 0.75 & 1 & 28.6312 & 198 & 5 & 27985743 & 11.113570 \\ 35 & 0.75 & 1 & 22.8631 & 191 & 6 & 22339313 & 10.276111 \\ 40 & 0.75 & 1 & 333.888 & 272 & 6 & 279093966 & 32.442125 \\ 40 & 0.75 & 1 & 309.575 & 264 & 9 & 261737608 & 149.204937 \\ 40 & 0.75 & 1 & 423.295 & 277 & 5 & 354920538 & 96.527282 \\ 40 & 0.75 & 1 & 255.759 & 261 & 10 & 209231563 & 153.626048 \\ 40 & 0.75 & 1 & 216.547 & 255 & 6 & 184462859 & 49.831940 \\ \hline \end{tabular} \end{center} \caption{Dense random graphs with edge probability $0.75$, $w\in[1,1000]$ and $w\in[1,1]$. Trivial lower bound.} \label{tab:dense-norebal} \end{table} \begin{table} \begin{center} \begin{tabular}{|rrr|r|r|rr|r|} \hline $n$ & Prob. & $\max w$ & Time & Cut & Solutions & Subproblems & Opt.\ Time \\ \hline 30 & 0.75 & 1000 & 0.103219 & 70916 & 11 & 96349 & 0.072895 \\ 30 & 0.75 & 1000 & 0.08406 & 68761 & 2 & 94414 & 0.016650 \\ 30 & 0.75 & 1000 & 0.042504 & 67895 & 2 & 47347 & 0.022612 \\ 30 & 0.75 & 1000 & 0.06844 & 66801 & 7 & 79773 & 0.058002 \\ 30 & 0.75 & 1000 & 0.072394 & 65323 & 11 & 82569 & 0.051860 \\ 35 & 0.75 & 1000 & 2.48431 & 101149 & 7 & 2724215 & 1.309536 \\ 35 & 0.75 & 1000 & 0.987641 & 88464 & 8 & 1011439 & 0.721924 \\ 35 & 0.75 & 1000 & 0.538253 & 91901 & 7 & 534528 & 0.519065 \\ 35 & 0.75 & 1000 & 0.58678 & 85501 & 11 & 574069 & 0.428499 \\ 35 & 0.75 & 1000 & 1.16092 & 87457 & 22 & 1185220 & 0.420064 \\ 40 & 0.75 & 1000 & 12.4906 & 131755 & 8 & 12185394 & 7.514862 \\ 40 & 0.75 & 1000 & 6.68102 & 121963 & 10 & 6317423 & 1.201540 \\ 40 & 0.75 & 1000 & 7.36096 & 126009 & 14 & 6943774 & 5.513143 \\ 40 & 0.75 & 1000 & 4.93249 & 119664 & 9 & 4624870 & 3.989565 \\ 40 & 0.75 & 1000 & 5.72612 & 114953 & 16 & 5422791 & 1.996018 \\ \hline 30 & 0.75 & 1 & 0.238 & 147 & 3 & 307003 & 0.191913 \\ 30 & 0.75 & 1 & 0.301815 & 150 & 4 & 379298 & 0.034334 \\ 30 & 0.75 & 1 & 0.189407 & 148 & 2 & 243934 & 0.181643 \\ 30 & 0.75 & 1 & 0.250688 & 148 & 1 & 316154 & 0.000008 \\ 30 & 0.75 & 1 & 0.258292 & 145 & 5 & 328164 & 0.063881 \\ 35 & 0.75 & 1 & 4.93283 & 205 & 1 & 5974035 & 0.000011 \\ 35 & 0.75 & 1 & 3.33951 & 195 & 4 & 3959624 & 1.939788 \\ 35 & 0.75 & 1 & 6.65941 & 206 & 6 & 8060167 & 4.803945 \\ 35 & 0.75 & 1 & 4.09707 & 198 & 5 & 4963393 & 1.857012 \\ 35 & 0.75 & 1 & 3.98347 & 191 & 6 & 4768656 & 1.942667 \\ 40 & 0.75 & 1 & 39.3355 & 272 & 6 & 43757388 & 3.069326 \\ 40 & 0.75 & 1 & 41.3252 & 264 & 9 & 46667399 & 20.465780 \\ 40 & 0.75 & 1 & 64.7553 & 277 & 5 & 72302392 & 14.290673 \\ 40 & 0.75 & 1 & 32.1253 & 261 & 10 & 34193672 & 18.884349 \\ 40 & 0.75 & 1 & 27.4485 & 255 & 6 & 29658475 & 5.525352 \\ \hline \end{tabular} \end{center} \caption{Dense random graphs with edge probability $0.75$, $w\in[1,1000]$ and $w\in[1,1]$. Lower bound with rebalancing contribution.} \label{tab:dense-rebal} \end{table} \begin{table} \begin{center} \begin{tabular}{|rrr|r|r|rr|r|} \hline $n$ & Prob. & $\max w$ & Time & Cut & Solutions & Subproblems & Opt.\ Time \\ \hline 30 & 0.75 & 1000 & 0.110197 & 70916 & 11 & 59014 & 0.073739 \\ 30 & 0.75 & 1000 & 0.110183 & 68761 & 2 & 57776 & 0.021733 \\ 30 & 0.75 & 1000 & 0.050028 & 67895 & 2 & 24208 & 0.026681 \\ 30 & 0.75 & 1000 & 0.093968 & 66801 & 7 & 49346 & 0.079876 \\ 30 & 0.75 & 1000 & 0.100189 & 65323 & 11 & 54630 & 0.070855 \\ 35 & 0.75 & 1000 & 3.26961 & 101149 & 7 & 1572739 & 1.722546 \\ 35 & 0.75 & 1000 & 1.4405 & 88464 & 8 & 642251 & 1.052328 \\ 35 & 0.75 & 1000 & 0.809674 & 91901 & 7 & 338742 & 0.782488 \\ 35 & 0.75 & 1000 & 0.845042 & 85501 & 11 & 357992 & 0.619300 \\ 35 & 0.75 & 1000 & 1.58746 & 87457 & 22 & 704004 & 0.571071 \\ 40 & 0.75 & 1000 & 14.2917 & 131755 & 8 & 5677755 & 8.680114 \\ 40 & 0.75 & 1000 & 10.558 & 121963 & 10 & 4217589 & 1.862208 \\ 40 & 0.75 & 1000 & 9.04176 & 126009 & 14 & 3509650 & 6.967428 \\ 40 & 0.75 & 1000 & 7.21843 & 119664 & 9 & 2803812 & 5.846905 \\ 40 & 0.75 & 1000 & 9.29349 & 114953 & 16 & 3709982 & 3.219045 \\ \hline 30 & 0.75 & 1 & 0.157407 & 147 & 3 & 92701 & 0.127816 \\ 30 & 0.75 & 1 & 0.192106 & 150 & 4 & 112422 & 0.022864 \\ 30 & 0.75 & 1 & 0.106529 & 148 & 2 & 60224 & 0.102453 \\ 30 & 0.75 & 1 & 0.146673 & 148 & 1 & 82837 & 0.000009 \\ 30 & 0.75 & 1 & 0.2163 & 145 & 5 & 128667 & 0.052001 \\ 35 & 0.75 & 1 & 2.55448 & 205 & 1 & 1294630 & 0.000010 \\ 35 & 0.75 & 1 & 1.64483 & 195 & 4 & 811987 & 1.005831 \\ 35 & 0.75 & 1 & 4.28534 & 206 & 6 & 2258574 & 3.174467 \\ 35 & 0.75 & 1 & 3.71848 & 198 & 5 & 1962477 & 1.797938 \\ 35 & 0.75 & 1 & 2.18208 & 191 & 6 & 1090292 & 1.109113 \\ 40 & 0.75 & 1 & 17.4109 & 272 & 6 & 7898119 & 1.560261 \\ 40 & 0.75 & 1 & 26.1942 & 264 & 9 & 12358406 & 13.616799 \\ 40 & 0.75 & 1 & 20.491 & 277 & 5 & 9320364 & 5.435163 \\ 40 & 0.75 & 1 & 16.4077 & 261 & 10 & 7301820 & 9.935822 \\ 40 & 0.75 & 1 & 18.6961 & 255 & 6 & 8532342 & 3.864666 \\ \hline \end{tabular} \end{center} \caption{Dense random graphs with edge probability $0.75$, $w\in[1,1000]$ and $w\in[1,1]$. Lower bound with rebalancing and high-degree contributions.} \label{tab:dense-highdegree} \end{table} \begin{table} \begin{center} \begin{tabular}{|rrr|r|r|rrr|r|} \hline $n$ & Prob. & $\max w$ & Time & Cut & Solutions & Subproblems & With optimal & Opt.\ Time \\ \hline 30 & 0.75 & 1000 & 0.111838 & 70916 & 11 & 59014 & 49411 & 0.074850 \\ 30 & 0.75 & 1000 & 0.111387 & 68761 & 2 & 57776 & 55645 & 0.022114 \\ 30 & 0.75 & 1000 & 0.050347 & 67895 & 2 & 24208 & 23955 & 0.026551 \\ 30 & 0.75 & 1000 & 0.095215 & 66801 & 7 & 49346 & 36983 & 0.080854 \\ 30 & 0.75 & 1000 & 0.101543 & 65323 & 11 & 54630 & 36903 & 0.071837 \\ 35 & 0.75 & 1000 & 3.28723 & 101149 & 7 & 1572739 & 1425159 & 1.734716 \\ 35 & 0.75 & 1000 & 1.44365 & 88464 & 8 & 642251 & 549589 & 1.059350 \\ 35 & 0.75 & 1000 & 0.803225 & 91901 & 7 & 338742 & 274701 & 0.776440 \\ 35 & 0.75 & 1000 & 0.849026 & 85501 & 11 & 357992 & 294054 & 0.623855 \\ 35 & 0.75 & 1000 & 1.58858 & 87457 & 22 & 704004 & 578057 & 0.566048 \\ 40 & 0.75 & 1000 & 14.177 & 131755 & 8 & 5677755 & 4348704 & 8.611801 \\ 40 & 0.75 & 1000 & 10.4853 & 121963 & 10 & 4217589 & 3966813 & 1.860220 \\ 40 & 0.75 & 1000 & 8.99509 & 126009 & 14 & 3509650 & 2261887 & 6.931919 \\ 40 & 0.75 & 1000 & 7.1699 & 119664 & 9 & 2803812 & 2452372 & 5.804552 \\ 40 & 0.75 & 1000 & 9.26082 & 114953 & 16 & 3709982 & 3174225 & 3.209197 \\ \hline 30 & 0.75 & 1 & 0.157288 & 147 & 3 & 92700 & 72481 & 0.127690 \\ 30 & 0.75 & 1 & 0.191835 & 150 & 4 & 112422 & 104662 & 0.022669 \\ 30 & 0.75 & 1 & 0.105953 & 148 & 2 & 60224 & 34917 & 0.101923 \\ 30 & 0.75 & 1 & 0.146986 & 148 & 1 & 82837 & 82837 & 0.000009 \\ 30 & 0.75 & 1 & 0.215565 & 145 & 5 & 128666 & 109322 & 0.051429 \\ 35 & 0.75 & 1 & 2.55151 & 205 & 1 & 1294624 & 1294624 & 0.000011 \\ 35 & 0.75 & 1 & 1.63822 & 195 & 4 & 811984 & 695869 & 1.002095 \\ 35 & 0.75 & 1 & 4.27547 & 206 & 6 & 2258569 & 1735175 & 3.168967 \\ 35 & 0.75 & 1 & 3.71503 & 198 & 5 & 1962473 & 1535044 & 1.798483 \\ 35 & 0.75 & 1 & 2.17589 & 191 & 6 & 1090289 & 916161 & 1.106713 \\ 40 & 0.75 & 1 & 17.4431 & 272 & 6 & 7898116 & 7616903 & 1.562321 \\ 40 & 0.75 & 1 & 26.1431 & 264 & 9 & 12358396 & 8928926 & 13.591346 \\ 40 & 0.75 & 1 & 20.3156 & 277 & 5 & 9320364 & 8395805 & 5.417039 \\ 40 & 0.75 & 1 & 16.2719 & 261 & 10 & 7301815 & 5643512 & 9.864531 \\ 40 & 0.75 & 1 & 18.6772 & 255 & 6 & 8532341 & 7852227 & 3.847519 \\ \hline \end{tabular} \end{center} \caption{Dense random graphs with edge probability $0.75$, $w\in[1,1000]$ and $w\in[1,1]$. Lower bound with rebalancing, high-degree and large connected component contributions.} \label{tab:dense-all} \end{table} The results for the five dense graphs are shown in in Table~\ref{tab:dense-norebal}, Table~\ref{tab:dense-rebal}, Table~\ref{tab:dense-highdegree} and Table~\ref{tab:dense-all}. As for the other graphs, the rebalancing contribution has the largest effect, and is huge. The high-degree bound now gives a significant reduction in number of subproblems, especially for the unweighted problems where the reduction is large enough to lead to a worthwhile reduction in running time. The large connected component contribution is rarely triggered here, leads only to a very small change in number of subproblems, and overall hardly affects the running time. The reduction in number of subproblems when the optimal solution is known initially is not as large as for the previous cases. \subsubsection{Complete graphs} \begin{table} \begin{center} \begin{tabular}{|rrr|r|r|rr|r|} \hline $n$ & Prob. & $\max w$ & Time & Cut & Solutions & Subproblems & Opt.\ Time \\ \hline 20 & 1 & 1000 & 0.006859 & 44780 & 2 & 5911 & 0.000485 \\ 20 & 1 & 1000 & 0.005944 & 40637 & 8 & 5308 & 0.001448 \\ 20 & 1 & 1000 & 0.006428 & 44723 & 2 & 5745 & 0.005717 \\ 20 & 1 & 1000 & 0.004846 & 41657 & 4 & 4204 & 0.001624 \\ 20 & 1 & 1000 & 0.005349 & 40891 & 9 & 4967 & 0.004198 \\ 30 & 1 & 1000 & 1.20566 & 99972 & 7 & 1074882 & 0.236397 \\ 30 & 1 & 1000 & 1.14248 & 91583 & 9 & 1007547 & 1.114057 \\ 30 & 1 & 1000 & 1.0778 & 96494 & 10 & 958071 & 0.500738 \\ 30 & 1 & 1000 & 1.08817 & 96948 & 4 & 967051 & 0.008571 \\ 30 & 1 & 1000 & 0.971279 & 93390 & 12 & 843119 & 0.814758 \\ \hline 20 & 1 & 1 & 0.01453 & 100 & 1 & 24309 & 0.000004 \\ 20 & 1 & 1 & 0.014483 & 100 & 1 & 24309 & 0.000003 \\ 20 & 1 & 1 & 0.014618 & 100 & 1 & 24309 & 0.000004 \\ 20 & 1 & 1 & 0.014633 & 100 & 1 & 24309 & 0.000003 \\ 20 & 1 & 1 & 0.014174 & 100 & 1 & 24309 & 0.000003 \\ 30 & 1 & 1 & 15.0829 & 225 & 1 & 20058299 & 0.000005 \\ 30 & 1 & 1 & 15.3418 & 225 & 1 & 20058299 & 0.000006 \\ 30 & 1 & 1 & 15.2103 & 225 & 1 & 20058299 & 0.000006 \\ 30 & 1 & 1 & 15.3602 & 225 & 1 & 20058299 & 0.000006 \\ 30 & 1 & 1 & 15.4689 & 225 & 1 & 20058299 & 0.000007 \\ \hline \end{tabular} \end{center} \caption{Complete random graphs, $w\in[1,1000]$ and $w\in[1,1]$. Trivial lower bound.} \label{tab:complete-norebal} \end{table} \begin{table} \begin{center} \begin{tabular}{|rrr|r|r|rr|r|} \hline $n$ & Prob. & $\max w$ & Time & Cut & Solutions & Subproblems & Opt.\ Time \\ \hline 20 & 1 & 1000 & 0.005071 & 44780 & 2 & 3754 & 0.000456 \\ 20 & 1 & 1000 & 0.003911 & 40637 & 8 & 3032 & 0.001306 \\ 20 & 1 & 1000 & 0.004503 & 44723 & 2 & 3580 & 0.004025 \\ 20 & 1 & 1000 & 0.002315 & 41657 & 4 & 1985 & 0.000856 \\ 20 & 1 & 1000 & 0.003182 & 40891 & 9 & 2663 & 0.002738 \\ 30 & 1 & 1000 & 0.335414 & 99972 & 7 & 356804 & 0.074178 \\ 30 & 1 & 1000 & 0.324173 & 91583 & 9 & 344801 & 0.317206 \\ 30 & 1 & 1000 & 0.259961 & 96494 & 10 & 283930 & 0.135704 \\ 30 & 1 & 1000 & 0.306426 & 96948 & 4 & 334555 & 0.004207 \\ 30 & 1 & 1000 & 0.241482 & 93390 & 12 & 251313 & 0.207057 \\ \hline 20 & 1 & 1 & 0.01766 & 100 & 1 & 24309 & 0.000005 \\ 20 & 1 & 1 & 0.017638 & 100 & 1 & 24309 & 0.000004 \\ 20 & 1 & 1 & 0.017909 & 100 & 1 & 24309 & 0.000004 \\ 20 & 1 & 1 & 0.017622 & 100 & 1 & 24309 & 0.000004 \\ 20 & 1 & 1 & 0.017693 & 100 & 1 & 24309 & 0.000003 \\ 30 & 1 & 1 & 17.8267 & 225 & 1 & 20058299 & 0.000008 \\ 30 & 1 & 1 & 17.9244 & 225 & 1 & 20058299 & 0.000007 \\ 30 & 1 & 1 & 17.8562 & 225 & 1 & 20058299 & 0.000007 \\ 30 & 1 & 1 & 18.5786 & 225 & 1 & 20058299 & 0.000008 \\ 30 & 1 & 1 & 17.7369 & 225 & 1 & 20058299 & 0.000008 \\ \hline \end{tabular} \end{center} \caption{Complete random graphs, $w\in[1,1000]$ and $w\in[1,1]$. Lower bound with rebalancing contribution.} \label{tab:complete-rebal} \end{table} \begin{table} \begin{center} \begin{tabular}{|rrr|r|r|rr|r|} \hline $n$ & Prob. & $\max w$ & Time & Cut & Solutions & Subproblems & Opt.\ Time \\ \hline 20 & 1 & 1000 & 0.001866 & 44780 & 2 & 1194 & 0.000202 \\ 20 & 1 & 1000 & 0.001798 & 40637 & 8 & 1154 & 0.000523 \\ 20 & 1 & 1000 & 0.001702 & 44723 & 2 & 1114 & 0.001546 \\ 20 & 1 & 1000 & 0.000833 & 41657 & 4 & 529 & 0.000307 \\ 20 & 1 & 1000 & 0.001809 & 40891 & 9 & 1250 & 0.001554 \\ 30 & 1 & 1000 & 0.157174 & 99972 & 7 & 79146 & 0.029180 \\ 30 & 1 & 1000 & 0.166044 & 91583 & 9 & 83938 & 0.162576 \\ 30 & 1 & 1000 & 0.117455 & 96494 & 10 & 58498 & 0.064518 \\ 30 & 1 & 1000 & 0.124764 & 96948 & 4 & 60501 & 0.001721 \\ 30 & 1 & 1000 & 0.114907 & 93390 & 12 & 56732 & 0.098666 \\ \hline 20 & 1 & 1 & 7e-06 & 100 & 1 & 0 & 0.000006 \\ 20 & 1 & 1 & 6e-06 & 100 & 1 & 0 & 0.000006 \\ 20 & 1 & 1 & 6e-06 & 100 & 1 & 0 & 0.000004 \\ 20 & 1 & 1 & 6e-06 & 100 & 1 & 0 & 0.000005 \\ 20 & 1 & 1 & 6e-06 & 100 & 1 & 0 & 0.000005 \\ 30 & 1 & 1 & 9e-06 & 225 & 1 & 0 & 0.000008 \\ 30 & 1 & 1 & 9e-06 & 225 & 1 & 0 & 0.000008 \\ 30 & 1 & 1 & 9e-06 & 225 & 1 & 0 & 0.000008 \\ 30 & 1 & 1 & 9e-06 & 225 & 1 & 0 & 0.000008 \\ 30 & 1 & 1 & 8e-06 & 225 & 1 & 0 & 0.000008 \\ \hline \end{tabular} \end{center} \caption{Complete random graphs, $w\in[1,1000]$ and $w\in[1,1]$. Lower bound with rebalancing and high-degree contributions.} \label{tab:complete-highdegree} \end{table} \begin{table} \begin{center} \begin{tabular}{|rrr|r|r|rrr|r|} \hline $n$ & Prob. & $\max w$ & Time & Cut & Solutions & Subproblems & With optimal & Opt.\ Time \\ \hline 20 & 1 & 1000 & 0.001868 & 44780 & 2 & 1194 & 1143 & 0.000206 \\ 20 & 1 & 1000 & 0.001751 & 40637 & 8 & 1154 & 1118 & 0.000518 \\ 20 & 1 & 1000 & 0.001712 & 44723 & 2 & 1114 & 866 & 0.001556 \\ 20 & 1 & 1000 & 0.000822 & 41657 & 4 & 529 & 471 & 0.000306 \\ 20 & 1 & 1000 & 0.001801 & 40891 & 9 & 1250 & 925 & 0.001548 \\ 30 & 1 & 1000 & 0.157706 & 99972 & 7 & 79146 & 76482 & 0.029056 \\ 30 & 1 & 1000 & 0.165645 & 91583 & 9 & 83938 & 66955 & 0.162109 \\ 30 & 1 & 1000 & 0.116554 & 96494 & 10 & 58498 & 41410 & 0.063725 \\ 30 & 1 & 1000 & 0.124496 & 96948 & 4 & 60501 & 60086 & 0.001730 \\ 30 & 1 & 1000 & 0.114147 & 93390 & 12 & 56732 & 44219 & 0.098063 \\ \hline 20 & 1 & 1 & 7e-06 & 100 & 1 & 0 & 0 & 0.000006 \\ 20 & 1 & 1 & 6e-06 & 100 & 1 & 0 & 0 & 0.000005 \\ 20 & 1 & 1 & 6e-06 & 100 & 1 & 0 & 0 & 0.000005 \\ 20 & 1 & 1 & 5e-06 & 100 & 1 & 0 & 0 & 0.000004 \\ 20 & 1 & 1 & 5e-06 & 100 & 1 & 0 & 0 & 0.000004 \\ 30 & 1 & 1 & 1e-05 & 225 & 1 & 0 & 0 & 0.000009 \\ 30 & 1 & 1 & 9e-06 & 225 & 1 & 0 & 0 & 0.000008 \\ 30 & 1 & 1 & 1e-05 & 225 & 1 & 0 & 0 & 0.000009 \\ 30 & 1 & 1 & 9e-06 & 225 & 1 & 0 & 0 & 0.000009 \\ 30 & 1 & 1 & 9e-06 & 225 & 1 & 0 & 0 & 0.000008 \\ \hline \end{tabular} \end{center} \caption{Complete random graphs, $w\in[1,1000]$ and $w\in[1,1]$. Lower bound with rebalancing, high-degree and large connected component contributions.} \label{tab:complete-all} \end{table} Results for the complete graphs can be found in Table~\ref{tab:complete-norebal}, Table~\ref{tab:complete-rebal}, Table~\ref{tab:complete-highdegree} and Table~\ref{tab:complete-all}. Here, the rebalancing contribution is much smaller, and only for weighted graphs; but as can be expected the high-degree contribution can instead be significant. Indeed, for the unweighted graphs, this leads to a bound which immediately proves that the initial, heuristic solution is optimal, and a reduction in number of subproblems from 20058299 to 0. The large connected components contribution is of course not triggered. \subsection{Parallel computing aspects} To illustrate that the Pheet framework can efficiently distribute the branch-and-bound search over a (large) number of cores, we include results for the parallel solution of some of the graph problems from the previous sections using now a prioritized search with some of the priority data structures implemented in Pheet. For details, see again~\cite{Wimmer14:diss}, and also~\cite{Traff13:priosched,Traff13:stratcorr}. The Pheet framework with the branch-and-bound code and the lower bounds developed in this report can be downloaded from \url{www.pheet.org}. The parallel graph partitioning experiments were performed on an 80-core Intel system with 1TB of memory consisting of eight 10-core Xeon E7-8850 processors. Experiments were run under Debian Linux and the framework compiled with \texttt{gcc 4.9.1}. The plots in Table~\ref{fig:sparse-60} to Table~\ref{fig:large-40} illustrate the speed-ups that can be achieved with increasing number of cores. The reported running times in seconds are the averages of 30 repeated runs with one graph type. \begin{figure} \includegraphics[width=\textwidth]{gcc_mars_40_10_1000_logic.pdf} \caption{Scalability for a sparse graph, $n=60$, edge probability $0.1$, $w\in [1,1000]$ with different scheduling strategies with 1 to 80 cores.} \label{fig:sparse-60} \end{figure} \begin{figure} \includegraphics[width=\textwidth]{gcc_mars_40_10_1000_logic.pdf} \caption{Scalability for a medium dense graph, $n=45$, edge probability $0.5$, $w\in [1,1000]$ with different scheduling strategies with 1 to 80 cores.} \label{fig:medium-45} \end{figure} \begin{figure} \includegraphics[width=\textwidth]{gcc_mars_40_10_1000_logic.pdf} \caption{Scalability for a dense graph, $n=40$, edge probability $0.75$, $w\in [1,1000]$ with different scheduling strategies with 1 to 80 cores.} \label{fig:large-40} \end{figure} Scheduling strategies make it possible to prioritize tasks representing graph partitioning subproblems, and select the most promising task for processing. Most promising can mean either the globally best task, the locally best task, or the task that is globally best according to a relaxed correctness criterion~\cite{Wimmer14:diss}. In the experiments a basic work-stealing scheduler (legend ``BasicScheduler'' and ``NoPriority'') not supporting priorities was compared against schedulers supporting strategies and priority queues with relaxed semantics (legend ``BStrategyScheduler'' and ``RelaxedPriority''). The rebalancing lower bound (legend ``Rebalancing'') is compared against the full bound with rebalancing, high-degree and connected-components contributions (legend ``Fullbound''). As can be seen in the three concrete cases, running times decreases with increasing number of cores, up till at least half the machine (40 cores). Prioritizing tasks provide significant reductions in running time. It is also interesting that the full bound, which in the sequential setting was often more expensive than the rebalancing bound becomes cheaper than the rebalancing bound as the number of cores increase (after four cores). \section{Concluding remarks} The purpose of this note was to resurrect and improve an old, combinatorial lower bound for the weighted graph partitioning problem, and to use this lower bound together with a modern, parallel task-scheduling framework for solving weighted graph bipartitioning problems as fast as possible. The results presented here a preliminary, and a number of possible improvements were discussed. The challenge to see whether the bound and the framework is competitive with current state-of-the art (combinatorial) approaches for the exact solution of graph partitioning problems (for certain types of graphs) remains. \bibliographystyle{abbrv}
\section{Introduction} In 1976 Chandrasekhar \cite{ChandraDIRACinKERRgeom,ChandraDIRACinKERRgeomERR}, Page \cite{Page76}, and Toop \cite{Too76} showed that Dirac's equation for a point electron in the Kerr--Newman spacetime \cite{NCCEPT65} separates essentially completely \footnote{In contrast to the familiar separation-of-variables results for, say, the Laplacian in a rectangular box or a cylinder or a sphere, Chandrasekhar, Page, and Toop obtained a system of ODEs for functions of only one variable each which is \emph{not} of triangular structure, and so cannot be solved one equation at a time.} in oblate spheroidal coordinates. Although this remarkable discovery enabled detailed mathematical studies of the behavior of a Dirac electron in a charged, rotating black hole spacetime \cite{KalMil1992,BelMar99,FinsterETalDperDNEerr,FinsterETalDperDNE,FinsterETalDKNa,FinsterETalDKNb,WINKLMEIERa,WINKLMEIERb,WINKLMEIERc,BelCac2010} (see also \cite{ChandraBOOKonBH} for neutral rotating black holes), there are perplexing conceptual issues which await clarification. Beside those that hark back to the enigmatic quantum-mechanical meaning of Dirac's equation in Minkowski spacetime, see \cite{Thaller} and \cite{KieTah14b}, serious new issues arise because of the physically somewhat questionable character of the Kerr--Newman solution, unveiled by Carter \cite{Car68}; see also \cite{ONeillBOOK,HehlREVIEW}. Namely, the maximal analytical extension of the stationary axisymmetric Kerr--Newman spacetime has a very strong curvature singularity on a timelike\footnote{In a limiting sense of course, since the metric is singular on this surface.} cylindrical surface whose cross-section with constant-$t$ hypersurfaces is a circle; here, $t$ is a coordinate pertinent to the \emph{asymptotically (at spacelike $\infty$) timelike} Killing field that encodes the stationarity of the ``outer regions'' of the Kerr--Newman spacetime. This circle is commonly referred to as \emph{the ``ring'' singularity}. The region near the ring is especially pathological since it harbors closed timelike loops.\footnote{The timelike ring singularity of the Kerr--Newman manifold is itself the limit of closed timelike loops, for which reason it is not possible to interpret this singular source of the stationary and axisymmetric Kerr--Newman electromagnetic fields outside of the outer ergosphere horizon as a ``rotating charged ring.'' For a careful analysis of the ring sources of the electromagnetic z$\,G$KN fields, see \cite{zGKN}.} Carter \cite{Car68} also showed that the maximal analytically extended Kerr--Newman manifold is ``cross-linked through the ring.''\footnote{The complement of a wedding ring in ordinary three-dimensional Euclidean space is topologically non-trivial, too, but ``looping through the ring once brings you back to where you began;'' in a spacelike slice of the maximal analytically extended Kerr--Newman spacetime ``you need to loop through the ring twice to get back to square one.''} Interestingly, this non-trivial topology was discovered a few years earlier in a family of static vacuum spacetimes by Zipoy \cite{Zip66}, who completely described their maximal analytical extension. Since Zipoy seems to have been the first to discover this non-trivial topology in exact spacetime solutions to Einstein's vacuum equations, we henceforth will refer to it as the \emph{Zipoy topology}. Carter showed that the Zipoy topology survives the vanishing-charge limit of the Kerr--Newman manifold, which yields the maximal analytic extension \cite{BoyLin67} of Kerr's solution \cite{Ker63} to Einstein's vacuum equations (${R}_{\mu\nu}=0$), cf.~\cite{HawEll73}. He furthermore showed that this topology also survives the vanishing-mass limit of the Kerr manifold, which yields an otherwise flat vacuum spacetime consisting of two static spacetime ends which are cross-linked through the ring. This vanishing-mass limit of the Kerr manifold coincides with the vanishing-mass limit of Zipoy's oblate spheroidal family of static vacuum spacetimes\footnote{In the same paper Zipoy also described another, prolate spheroidal family of static vacuum spacetimes, whose metric is nowadays known as ``Zipoy--Voorhees metric."} In the black-hole sector of their parameter space the Kerr--Newman spacetimes also have a Cauchy horizon, an event horizon, and an ergosphere horizon; see \cite{HawEll73,ONeillBOOK,HehlREVIEW}. From the ``safe perspective of an observer at spatial infinity'' the ring singularity, the acausal region, and the Cauchy horizon are invisible, being ``hidden'' behind the event horizon, and no exotic or even objectionable physics would ever seem to happen: a Dirac spinor wavefunction initially supported outside the event horizon will either keep spreading within the outer region or eventually (as $t\to\infty$) accumulate (in parts or wholly) at the event horizon, see \cite{FinsterETalDKNa,FinsterETalDKNb}. Yet, an inquisitive physicist may also want to study the spinor wave function in other coordinates designed to ``follow it across the event horizon,''\footnote{When the analogous study was carried out by Oppenheimer and Snyder \cite{OppSny} for classical gaseous matter undergoing gravitational collapse it leveled the ground for building our modern understanding of the physics of gravitational collapse, involving the formation of black holes and their singularities. Poetically speaking their work revealed that there is more physics in general relativity than meets the (distant observer's) eye.} however, it is neither clear how to continue in a ``physically correct'' manner beyond the Cauchy horizon\footnote{If instead of the Cauchy problem one studies $t$-periodic solutions, then one can continue across the Cauchy and the event horizons using a weak matching procedure \cite{FinsterETalDperDNEerr,FinsterETalDperDNE}. However, Finster et al. \cite{FinsterETalDperDNEerr,FinsterETalDperDNE} found that no $t$-periodic solutions exist which are normalized over a constant-$t$ slice of the ``physical black hole spacetime;'' see main text.} nor what to make of the acausal region of closed timelike loops, nor how to correctly handle the timelike singularity. Moreover, if one inquires into the ``physics beyond the event horizon,'' one has the option of allowing the support of the initial spinor wave function to be spread over both asymptotically flat ends, or some other parts of the maximal analytically extended Kerr--Newman spacetime. It is not so clear which options (if any) are physically reasonable and which ones are science fiction, although astrophysicists can argue for the ``physical black hole spacetime,'' i.e. the part of the maximal analytically extended Kerr--Newman spacetime which is the asymptotic limit of the topologically simple spacetime of a charged, rotating star collapsing into a black hole \cite{StraumannBOOK}. The horizons are absent in the hyper-extremal parameter regime. Even though the absence of the Cauchy horizon is a welcome simplification, this regime is rarely studied because the absence of the event horizon renders the singularity ``naked,'' and the (weak) cosmic censorship hypothesis, according to which ``nature abhors naked singularities'' \cite{PenroseNAKEDsing,PenroseCOSMICcensor}, has (unfortunately) discouraged physicists from investigating spacetimes with naked singularities. Yet once a piece of a Dirac spinor wave function has crossed the event horizon of a Kerr--Newman black hole it is no longer shielded from possible harm done by the spacetime singularity --- viz., inside the event horizon the singularity \emph{is} naked ---, and so one may as well study the effects of naked singularities directly. Be that as it may, the hyper-extremal regime retains the closed timelike loops which according to the standard interpretation of general relativity turn the entire manifold into a \emph{causally vicious set}, something that many physicists (including the authors) would regard as physically suspicious. A strategy to rid the Kerr--Newman manifold from its Cauchy horizon, and all its other acausal aspects, is to take a \emph{zero-gravity limit} $G\to 0$, where $G$ is Newton's constant of universal gravitation. This would be quite uninteresting if the zero-$G$ limit of the Kerr--Newman manifold would simply yield a Minkowski spacetime decorated with the electric field of a point charge and the magnetic field of a point dipole, as one might be tempted to guess from the asymptotically flat ends of the Kerr--Newman spacetime. However, as shown by one of us in the accompanying paper \cite{zGKN}, the zero-gravity limit of the maximal analytically extended Kerr--Newman spacetime yields a static, flat, yet \emph{two-leafed, cross-linked} spacetime which is decorated with Appell--Sommerfeld \cite{Appell,Som97} electromagnetic fields\footnote{The zero-$G$ limit of the electromagnetic Kerr--Newman fields yields fields originally discovered by Appell \cite{Appell}, who obtained them from the Coulomb potential of a point charge by a complex translation of the charge's position. Appell noticed that the fields change sign when looping once through the ring, but did not conclude --- apparently --- that they live naturally on a topologically nontrivial space. Sommerfeld \cite{Som97} seems to have been first to introduce ``branched Riemann spaces,'' three-dimensional analogues of topologically non-trivial Riemann surfaces, to which we will refer as \emph{Sommerfeld spaces}, and to construct electromagnetic fields (harmonic functions) on them, which in general we will call \emph{Sommerfeld fields}. Eventually Evans \cite{Evans51} and his students \cite{Neu51}, \cite{Alz04} laid their rigorous foundations.} $\mathbf{F} = d\mathbf{A}$ with $\mathbf{A}$ as in \refeq{def:AKN}, whose sources are certain ``finite charge and current distributions'' supported by the one-dimensional ring.\footnote{Strictly speaking, the ring singularity is not part of the manifold; it's rather a ring ``defect.''} Although the gravitational (viz.: curvature) aspects of the Kerr--Newman manifold, its event horizon included, vanish in this limit too, one does retain the topological, the singular, and all the electromagnetic aspects of the spacetime. Studies of the Dirac equation for a point electron in this zero-$G$ Kerr--Newman (z$G$KN) spacetime will therefore illuminate the role played by the topological and electromagnetic aspects of the Kerr--Newman manifold in the relativistic quantum mechanics of the electron. In this paper we study the Dirac equation for a point electron in static, electromagnetic, flat spacetimes with Zipoy topology which include the z$G$KN spacetimes as special case, but which in general can sport Sommerfeld fields $\mathbf{F} = d\mathbf{A}$ with $\mathbf{A}$ as in \refeq{def:AQI}, which differ from the Appell--Sommerfeld fields only in a single number, the ratio $I\pi a^2/Qa$ of their magnetic dipole moment to the magnetic dipole moment of the Appell--Sommerfeld fields of same charge $Q$; here, $|a|$ is the radius of the ring singularity. By constructing an operator that anti-commutes with the pertinent Dirac Hamiltonian we show that the spectrum of any of its self-adjoint extensions is symmetric about zero; this result holds for arbitrary $(Q,I)$. All other results are obtained for Dirac's electron in the z$G$KN spacetime ($I\pi a^2/Qa=1$): by adapting an argument of Winklmeier--Yamada for the Dirac equation of a point electron in the outer region of the Kerr--Newman black hole spacetime, we show that our formal Dirac Hamiltonian is essentially self-adjoint on a spacelike slice of the maximal analytically extended, static z$G$KN spacetime. Then we exploit the Chandrasekhar--Page--Toop separation-of-variables method for Dirac's equation on a general Kerr--Newman spacetime, and the Pr\"ufer transform, to show that the self-adjoint Dirac operator has a continuous spectrum with a gap about zero which, under two smallness conditions, contains a pure point spectrum associated with time-periodic $L^2$ spinor fields, representing bound states of Dirac's point electron in the electromagnetic field of the ring singularity of the z$G$KN spacetime. In the next section we formulate our main results about the Dirac equation for a point electron in the z$G$KN spacetime; one result is valid also for a Dirac electron in static, flat spacetimes having Zipoy topology featuring electromagnetic Sommerfeld fields of arbitrary $I\pi a/Q$-ratio. In sections 3, 4, 5, 6, and 7 we prove our main theorems about the spectrum. In section 8 we conclude with a list of interesting questions left unanswered by this work. \vskip-0.9truecm $\phantom{nix}$ \section{Formulation of the main results} We begin by formulating the Dirac equation for a point electron in electromagnetic, static, flat spacetimes with Zipoy topology which generalize zero-$G$ Kerr--Newman spacetimes to zero-$G$ Kerr spacetimes equipped with Sommerfeld fields of arbitrary $I\pi a/Q$-ratio. We then state our main theorems about the spectrum of the pertinent Dirac operators. \vskip-20pt$\phantom{nix}$ \subsection{Dirac's equation for a point electron on zero-$G$ Kerr spacetimes equipped with electromagnetic Sommerfeld fields of arbitrary $I\pi a/Q$-ratio} \subsubsection{Zero-$G$ Kerr spacetimes}\label{sec:zGK} Our limit $G\to 0$ of the maximal analytic extension of the well-known {\em Kerr} family of stationary, axisymmetric spacetime solutions of Einstein's vacuum equations yields a one-parameter family of static, flat, but topologically nontrivial spacetimes\footnote{ We emphasize that Zipoy \cite{Zip66} found a large class of static, axisymmetric, flat but topologically nontrivial solutions to the Einstein vacuum equations which in their zero-$G$ limit coincide with the members of the zero-$G$ Kerr spacetime family.} $({\mathcal M},\mathbf{g})$ which consist of two ``cross-linked leafs.'' Explicitly, let $\mathcal{C}\equiv\{(t,r,\theta,\varphi): t\in{\mathbb R}, r\in{\mathbb R},\theta\in[0,\pi], \varphi\in[0,2\pi)\}$ denote a rectangular ``four-dimensional cylinder,'' and let ${\mathcal S}\equiv \{(t,r,\theta,\varphi): t\in{\mathbb R}, r=0, \theta =\pi/2,\varphi\in[0,2\pi)\}\subset\mathcal{C}$ denote a rectangular ``two-dimensional slab'' in $\mathcal{C}$. Then $\mathcal{C}\setminus {\mathcal S}$ is a covering chart of oblate spheroidal (Boyer--Lindquist, or BL) coordinates\footnote{The notation $(t,r,\theta,\varphi)$ for the BL coordinates is standard in the relativity literature, and should not be confused for instance with the Schwarzschild coordinates on the outer region of that spacetime, or with just standard spherical coordinates of Minkowski spacetime. All standard non-flat $(t,r,\theta,\varphi)$ coordinate systems reduce to standard spherical coordinates of the flat Minkowski spacetime near ``$r = \infty$.'' Note that in BL coordinates $r$ takes any real value.} for this spacetime, with line element \begin{equation}\label{metricOS} ds_{\mathbf{g}}^2 = dt^2 - \left(r^2+a^2\right) \sin^2\theta\, d\varphi^2 - \frac{r^2+a^2 \cos^2\theta}{r^2+a^2} \left(dr^2 + \left(r^2+a^2\right) d\theta^2\right); \end{equation} here, $a^2>0$ is the only parameter of these spacetimes, and we have set the speed of light $c=1$. Our sign convention of $(+,-,-,-)$ for the metric follows \cite{FinsterETalDperDNE}. The static, axisymmetric character of these zero-$G$ Kerr spacetimes is manifest in (\ref{metricOS}). Also, since $r\in{\mathbb R}$ occurs strictly quadratically in (\ref{metricOS}), it is clear that the manifold consists of two ``conjoined identical twins.'' To exhibit their flatness, and in the process also the topological nontrivial juncture, we introduce cylindrical coordinates $(t,\varrho,z,\varphi)$ on Minkowski spacetime ${\mathbb R}^{1,3}$, with the same $(t,\varphi)$ as in Boyer--Lindquist coordinates, and with the cylindrical coordinates $(\varrho,z)$ related to the elliptical coordinates $(r,\theta)$ by \begin{equation}\label{CYLtoELL} \varrho = \sqrt{r^2+a^2}\sin\theta,\qquad z = r \cos \theta. \end{equation} In cylindrical $(t,\varrho,z,\varphi)$ coordinates the metric takes the familiar form for flat Minkowski spacetime \begin{equation} ds_{\mathbf{g}}^2 = dt^2 - d\varrho^2 - \varrho^2d\varphi^2 - dz^2, \end{equation} except that the map (\ref{CYLtoELL}) makes it plain that the chart $\mathcal{C}\setminus{\mathcal S}$ will be mapped into \emph{two} copies of Minkowski spacetime which are ``doubly conjoined,'' in a smooth yet crossing manner, at the set spanned by ${\mathcal S}$. The metric ${\mathbf{g}}$ given by the line element (\ref{metricOS}) has a singularity at ${\mathcal S}$, which is the singularity \emph{of the spacetime}, not \emph{in the spacetime}. The set ${\mathcal S}$ is the boundary of a timelike open solid cylinder in the z$G$KN manifold; the cross section at any instant $t$ of this cylindrical surface is a translate of the {\em ring} $\mathcal{R}_{0}\equiv\{(t,r,\theta,\varphi):t=0,r=0, \theta=\pi/2,\varphi\in[0,2\pi)\}$ of Euclidean radius $\sqrt{a^2}=|a|>0$, for which reason one speaks of a ``ring singularity;'' we write $|a|$ because we need to allow $a\in{\mathbb R}\backslash\{0\}$ for reasons that become clear in the next subsection. The points on the ring are {\em conical singularities} for the metric, meaning that the limit as the radius goes to zero of the ratio of the circumference to radius of a small circle centered at a point of the ring and lying in a meridional plane $\varphi=$const. is not $2\pi$; instead, here it is $4\pi$. See \cite{zGKN} for details. The key topological features of this manifold can easily be visualized. Namely, although the fixed timelike planes $\{(t,r,\theta,\varphi): t=t_0,\varphi=\varphi_0\}$ cannot be embedded into ${\mathbb R}^3$, each such plane can be immersed in it, the immersion consisting of two Euclidean half planes, stacked up upon each other, then cut along a line segment of length $|a|$ orthogonal to the planes' boundaries ``with scissors,'' then smoothly ``cross-glued'' at the cut such that the ``upper'' and ``lower'' sheets are cross-linked like an $\times$ along the cut, while remaining like $\|$ beyond the cut; the singular endpoint of the $\times$-line is not part of the two-sheeted manifold. Shown in Fig.1 are the ring singularity and the part $\{r\in(-1,1),\theta\in(0,\pi)\}$ of a constant-azimuth section (slightly curved, for the purpose of visualization) of the two-sheeted static spacelike slice of the zero-$G$ Kerr spacetime. The coordinate grid shows a few of the curves $\theta = \mbox{const.}$ (hyperbolas, transiting through the ring from one sheet to the other) and $r = \mbox{const.}$ (oblate semi-ellipses, remaining outside the ring on a single sheet). \vspace{-.1cm}$\phantom{nix}$ \begin{figure}[ht] \begin{center} \includegraphics[width=9cm,height=7cm]{ZIPOYtopologyTILTED} \end{center} \vspace{-2cm} \caption{An illustration of the Zipoy topology.\label{fig:ZipTop}} \end{figure} We can view ${\mathcal M}$ as a {\em branched covering} over the base manifold ${\mathbb R}^{1,3}\setminus{\mathcal S}$ (mildly abusing notation), with the projection map $\Pi: {\mathcal M} \to {\mathbb R}^{1,3}\setminus{\mathcal S}$ being $\Pi(t,r,\theta,\varphi) = (t,\varrho,z,\varphi)$. The pre-image of a point in the base consists of two points, degenerating into one point at each ``ring'' \begin{equation} \mathcal{R}_{t_0} = \{(t,r,\theta,\varphi)\ |\ t=t_0,\ r=0, \ \theta = \pi/2, 0\leq \varphi\ \leq 2\pi \}. \end{equation} The pullback of the Minkowski metric $\boldsymbol{\eta}$ under $\Pi$ endows ${\mathcal M}$ with a flat Lorentzian metric $\mathbf{g} = \Pi^*\boldsymbol{\eta}$, whose line element is given in (\ref{metricOS}). \subsubsection{Zero-G Kerr-Newman spacetimes}\label{sec:zGKN} The spacetime $({\mathcal M},\mathbf{g})$ introduced above can be decorated with any static electromagnetic Sommerfeld field $\mathbf{F}=d\mathbf{A}$, satisfying the flat space Maxwell equations locally but respecting the topologically nontrivial character of the spacetime. The zero-$G$ limit of the maximal analytical extension of the {\em Kerr--Newman} family\footnote{We recall that for fixed $G>0$ (and speed of light $c$) the Kerr--Newman family is a three-parameter family of electrovac spacetimes, the parameters being ADM mass (energy) $M(>0)$, ADM angular momentum $J=Ma\in{\mathbb R}$, and total charge $Q\in{\mathbb R}$, all defined in a single asymptotic end. Note that in units where $c=1$ the angular momentum per unit mass $a$ has physical dimension of length; equivalently, of time.} of stationary axisymmetric solutions to the Einstein--Maxwell equations, written in BL coordinates, yields precisely the zero-$G$ Kerr (z$G$K) spacetime decorated with a particular electromagnetic Sommerfeld field, the \emph{Appell--Sommerfeld field}, whose four-potential one-form reads \begin{equation}\label{def:AKN} \mathbf{A} = - \frac{r}{r^2+a^2\cos^2\theta} (Qdt - Qa \sin^2\theta\, d\varphi). \end{equation} Here, $Q$ is the total charge ``seen from infinity'' in the $r>0$ sheet, defined by computing the electric outward flux through a spherical surface surrounding the ring singularity in that sheet, and we need to allow $a\in{\mathbb R}\backslash\{0\}$ to accomodate a magnetic dipole moment vector pointing either ``up'' or ``down'' w.r.t. the $z$ axis defined in (\ref{CYLtoELL}). The field $\mathbf{F}$ is singular on the same ring $\mathcal{R}_t$ as is the metric, while for $r$ very large positive its electric and magnetic components approach, respectively, the asymptotics of an ``electric monopole field in ${\mathbb R}^3$ of a charge $Q$'' and a ``magnetic dipole field in ${\mathbb R}^3$ of dipole moment $Qa$,'' aligned parallel or anti-parallel to the ``$z$ direction;'' for $-r$ very large positive, i.e. in the other sheet, these electric monopole and magnetic dipole fields correspond to a charge $-Q$ and magnetic dipole moment $-Qa$. \begin{rem}\textit{ Note that in contrast to the metric (\ref{metricOS}), the electromagnetic potential (\ref{def:AKN}) is not invariant under a change of sign of $a\in {\mathbb R}\backslash\{0\}$; however, the difference is merely in the direction of the magnetic moment vector that corresponds to the above potential, which either points along or opposite to the ``$z$ direction'' defined by (\ref{CYLtoELL}). Since a 180-degree rotation around a diametrical axis through the ring transforms its magnetic moment into the negative thereof, and since all objectively physical quantities are invariant under such a rotation, the choices $a>0$ and $a<0$ are physically equivalent, given $Q$ and $a^2>0$. Therefore, without loss of generality, we can choose $a>0$ in all ensuing calculations involving (\ref{def:AKN}).} \end{rem} \begin{rem}\textit{ It may be tempting to speculate whether the magnetic dipole moment $Qa$ can be interpreted as due to a ``gyrating charged ring,'' with ``$a$'' the angular momentum per unit mass ``of the singularity,'' as has been attempted for Kerr--Newman spacetimes. Moreover, since Kerr--Newman spacetimes have a {\em gyromagnetic ratio} $Q/M =:g_{\mbox{\tiny{KN}}}^{}Q/2M$ (in units with $c=1$), amounting to a $g$-factor $g_{\mbox{\tiny{KN}}}^{}=2$ (see \cite{Car68}), and since the KN parameters $(M,Q,a)$ are independent of $G$, and so is the KN gyromagnetic ratio, one could be tempted to assign the z$\,G$KN spacetime the same gyromagnetic ratio of $Q/M$ and $g$-factor of $2$. However, since $M$ does not show in the z$\,G$KN metric, such an assignment would be reasonable only if there were no other way to construct z$\,G$KN than taking the zero-$G$ limit of KN. Yet this is not the case: as already pointed out in footnote 10, the underlying spacetime manifold of z$\,G$KN can be obtained as zero-$G$ limit of either, the \emph{stationary} family of Kerr spacetimes --- having both an ADM mass $M_{\mbox{\tiny{K}}}^{}$ and ADM angular momentum $J=M_{\mbox{\tiny{K}}}^{}a$ ---, or a \emph{static} family of Zipoy spacetimes --- having an ADM mass $M_{\mbox{\tiny{Z}}}^{}$ but zero ADM angular momentum; note also that $M_{\mbox{\tiny{Z}}}^{}\neq M_{\mbox{\tiny{K}}}^{}$ in general. This (say) z$\,G$Z spacetime can now be equipped with an arbitrary Sommerfeld field, in particular: the Appell--Sommerfeld field of z$\,G$KN, without being logical compelled to interpret its magnetic moment $Qa$ as being due to a ``gyrating ring of charge $Q$'' with angular momentum per unit mass ``$a$,'' although this is logically possible; in any event, it's better to refrain from assigning z$\,G$KN any {\em spacetime $g$-factor}. } \end{rem} \subsubsection{Generalizations of z$G$KN spacetimes to arbitrary charge and current.} Since the electric and magnetic components of Maxwell's vacuum equations decouple in the zero-$G$ limit, to decorate the zero-$G$ Kerr spacetime with a generalization of the electromagnetic Appell--Sommerfeld field $\mathbf{F} = d\mathbf{A}$ having electric charge $Q\in{\mathbb R}$ and current $I\in{\mathbb R}$, all that needs to be done is to replace the magnetic dipole moment $Qa$ in formula (\ref{def:AKN}) by $I\pi a^2$, thus \begin{equation}\label{def:AQI} \mathbf{A} = - \frac{r}{r^2+a^2\cos^2\theta} (Q dt - I\pi a^2 \sin^2\theta\, d\varphi). \end{equation} Again, electric charge $Q$ and magnetic dipole moment $I\pi a^2$ are as ``seen'' from spacelike infinity in the $r>0$ sheet; viewing from spacelike infinity in the other sheet one ``sees'' $-Q$ and $-I\pi a^2$. \begin{rem}\textit{ The Sommerfeld field (\ref{def:AQI}) makes it plain that the z$\,G$KN spacetimes are but a special one-parameter subfamily in a two-parameter family of qualitatively similar electromagnetic spacetimes with arbitrary charge $Q$ and magnetic moment $I\pi a^2$ (given $a^2$). The ease with which this result was accomplished stands in stark contrast to the difficulties in generalizing the Kerr--Newman family to electromagnetic spacetimes with a magnetic moment different from $Qa$. } \end{rem} \begin{rem}\textit{ As does the metric (\ref{metricOS}), the generalized Sommerfeld field (\ref{def:AQI}) depends on ``$a$'' only through $a^2$, hence it is invariant under a sign change of $a\in {\mathbb R}\backslash\{0\}$. So when (\ref{metricOS}) is combined with (\ref{def:AQI}) we can choose $a>0$ without any further ado. Of course, in (\ref{def:AQI}) there is now the new parameter $I\in{\mathbb R}$, but by the same rotation argument as given for the z$\,G$KN fields the choices $I>0$ and $I<0$ are physically equivalent, given $Q$ and $a^2>0$. Therefore, without loss of generality, from here on we choose ${\mathrm{sign\,}}(I)={\mathrm{sign\,}}(Q)$, and $a>0$.} \end{rem} \subsubsection{The Dirac equation on electromagnetic spacetimes: Cartan's frame method} In arbitrary cordinates $(x^\mu)$ (with $c=1$ and $\hbar=1$), the Dirac equation for a spin-$1/2$ electron of empirical rest mass $m$ and charge $-e<0$ interacting (through minimal coupling) with an electromagnetic field $\mathbf{F} = d\mathbf{A}$ in a spacetime $({\mathcal M},\mathbf{g})$ reads \begin{equation}\label{eq:DirEqA} {\tilde\gamma}^\mu (-i \nabla_\mu + e A_\mu) \Psi + m \Psi = 0; \end{equation} here $\nabla$ is the covariant derivative (on bi-spinors) associated to the spacetime metric $\mathbf{g}$, and the $({\tilde\gamma^\mu})_{\mu=0}^3$ are Dirac matrices associated to this metric, i.e. satisfying \begin{equation} \tilde{\gamma}^\mu \tilde{\gamma}^\nu + \tilde{\gamma}^\nu \tilde{\gamma}^\mu = 2 g^{\mu\nu}{\boldsymbol{1}}_{4\times4}, \end{equation} while the $A_\mu$ are the pertinent components of the electromagnetic potential, $\mathbf{A} = A_\mu dx^\mu$. Using Cartan's frame method (see \cite{BrillCohen66} and refs. therein) one can express the above covariant derivative on spinors in terms of standard derivatives: \begin{equation} \tilde{\gamma}^\mu \nabla_\mu = \gamma^\mu\mathbf{e}_\mu +\frac{1}{4} \Omega_{\mu\nu\lambda} \gamma^\lambda\gamma^\mu\gamma^\nu, \end{equation} where the $\gamma^\nu$ are Dirac gamma matrices for the Minkowski spacetime, satisfying \begin{equation}\label{MinkGamma} \gamma^\nu\gamma^\mu+\gamma^\mu\gamma^\nu = 2{{\eta}}^{\mu\nu}{\boldsymbol{1}}_{4\times4}, \end{equation} with \begin{equation} ({\boldsymbol{\eta}}) = \mbox{diag}(1,-1,-1,-1) \end{equation} being the matrix of the Minkowski metric in rectangular coordinates; and $\{\mathbf{e}_\mu\}_{\mu=0}^3$ is a {\em Cartan frame}, i.e. an orthonormal frame of vectors spanning the tangent space at each point of the spacetime manifold. We thus have \begin{equation} (\mathbf{e}_\mu)^\nu (\mathbf{e}_\lambda)^\kappa {g}_{\nu\kappa} = {{\eta}}_{\mu\lambda}. \end{equation} On the one hand, it follows that \begin{equation} \tilde{\gamma}^\mu = (\mathbf{e}_\nu)^\mu \gamma^\nu. \end{equation} On the other hand, let $\{\boldsymbol{\omega}^\mu\}_{\mu=0}^3$ denote the {\em dual} frame to $\{\mathbf{e}_\mu\}$, i.e. the orthonormal basis for the cotangent space at each point of the manifold that is dual to the basis for the tangent space: \begin{equation} \boldsymbol{\omega}_\mu \big(\mathbf{e}^\nu\big) = \mathbf{e}^\nu \big( \boldsymbol{\omega}_\mu\big) = \delta_\mu^\nu. \end{equation} Then the $\Omega_{\mu\nu\lambda}$ are by definition the {\em Ricci rotation coefficients} of the frame $\{\boldsymbol{\omega}^\mu\}_{\mu=0}^3$, defined in the following way: Let the one-forms $\boldsymbol{\Omega}^\mu_\nu$ satisfy \begin{equation} d \boldsymbol{\omega}^\mu + \boldsymbol{\Omega}^\mu_\nu \wedge \boldsymbol{\omega}^\nu = 0. \end{equation} This does not uniquely define the $\boldsymbol{\Omega}^\mu_\nu$. However, there exists a unique set of such 1-forms satisfying the extra condition \begin{equation} \boldsymbol{\Omega}_{\mu\nu} = - \boldsymbol{\Omega}_{\nu\mu}, \end{equation} where the first index is lowered by the Minkowski metric: $\boldsymbol{\Omega}_{\mu\nu} := {{\eta}}_{\mu\lambda}\boldsymbol{\Omega}^{\lambda}_\nu$. Since $\left\{\boldsymbol{\omega}^\mu\right\}$ forms a basis for the space of 1-forms, we then have $\boldsymbol{\Omega}_{\mu\nu} = \Omega_{\mu\nu\lambda}\boldsymbol{\omega}^\lambda$, which defines the rotation coefficients $\Omega_{\mu\nu\lambda}$. The Dirac equation (\ref{eq:DirEqA}) on a spacetime $({\mathcal M},\mathbf{g})$ with an electromagnetic 4-potential $\mathbf{A}$ can thus be written in the following form: \begin{equation}\label{eq:DirEqAstandard} \gamma^\mu \left(\mathbf{e}_\mu + \Gamma_\mu + i e \tilde{A}_\mu\right)\Psi + im\Psi = 0; \end{equation} here, the $\Gamma_\mu$ are connection coefficients, \begin{equation} \Gamma_\mu := \frac{1}{4} \Omega_{\nu\lambda\mu} \gamma^\nu\gamma^\lambda = \frac{1}{8} \Omega_{\nu\lambda\mu}[\gamma^\nu,\gamma^\lambda], \end{equation} and the $\tilde{A}_\mu$ are the components of the potential $\mathbf{A}$ in the ${\boldsymbol{\omega}^\mu}$ basis, i.e. $\mathbf{A} =\tilde{A}_\mu \boldsymbol{\omega}^\mu$, or, \begin{equation} \tilde{A}_\mu := (\mathbf{e}_\mu)^\nu A_\nu. \end{equation} \subsubsection{Frame formulation of the Dirac equation on z$G$K spacetimes featuring generalized electromagnetic Sommerfeld fields with arbitrary $I\pi a/Q$-ratio} As explained in section \ref{sec:zGK}, the single chart $\mathcal{C}\setminus{\mathcal S}$ of oblate spheroidal coordinates $(t,r,\theta,\varphi)$ covers the whole zero-$G$ Kerr spacetime $({\mathcal M},\mathbf{g})$, and in section \ref{sec:zGKN} we saw that in these coordinates the generalized electromagnetic Sommerfeld one-form $\mathbf{A}$ is everywhere on $({\mathcal M},\mathbf{g})$ given by the simple formula (\ref{def:AQI}). It is therefore natural that one would like to write Dirac's equation \refeq{eq:DirEqA} in these coordinates as well, in the hope of achieving at least some partial separation of variables.\footnote{The idea of using special frames adapted to a coordinate system in order to separate spinorial wave equations in those coordinates goes back to Kinnersley \cite{Kin69} and Teukolsky \cite{Teu72}.} However, unlike Cartesian coordinates $(x^\mu)$ in Minkowski spacetime, oblate spheroidal coordinate derivatives do not give rise to an orthonormal basis for the tangent space at each point of a zero-$G$ Kerr spacetime. Thus, to bring \refeq{eq:DirEqA} into the Cartan form \refeq{eq:DirEqAstandard} using oblate spheroidal coordinates, one also needs to construct a suitable Cartan frame. Following Chandrasekhar \cite{ChandraDIRACinKERRgeom,ChandraDIRACinKERRgeomERR}, Page \cite{Page76}, Toop \cite{Too76} (see also Carter-McLenaghan \cite{CarMcL82}), we introduce a special orthonormal frame $\{\mathbf{e}_\mu\}_{\mu=0}^3$ on the tangent bundle $T{\mathcal M}$ which is adapted to the oblate spheroidal coordinates, such that the Dirac equation takes a comparatively simple form. We begin by introducing a Cartan (co-)frame $\{\boldsymbol{\omega}^\mu\}_{\mu=0}^3$ for the cotangent bundle\footnote{This particular frame is called a {\em canonical symmetric tetrad} in \cite{CarMcL82}.} \begin{equation}\label{CcoF} \boldsymbol{\omega}^0 := \frac{\varpi}{|\rho|} (dt - a \sin^2\theta\, d\varphi),\quad \boldsymbol{\omega}^1 := |\rho|d\theta,\quad \boldsymbol{\omega}^2 := \frac{\sin\theta}{|\rho|} (-a dt + \varpi^2 d\varphi),\quad \boldsymbol{\omega}^3 := \frac{|\rho|}{\varpi}dr, \end{equation} with the abbreviations \begin{equation}\label{vpirho} \varpi := \sqrt{r^2 + a^2}, \quad \rho:= r + i a \cos\theta. \end{equation} Let us denote the oblate spheroidal coordinates $(t,r,\theta,\varphi)$ collectively by $(y^\nu)$. Let $g_{\mu\nu}$ denote the coefficients of the spacetime metric \refeq{metricOS} in oblate spheroidal coordinates, i.e. $g_{\mu\nu} = \mathbf{g}\Big(\frac{\partial}{\partial y^\mu},\frac{\partial}{\partial y^\nu}\Big)$. One easily checks that written in the $\{\boldsymbol{\omega}^{\mu}\}$ frame, the spacetime line element is \begin{equation} ds_{\mathbf{g}}^2 = g_{\mu\nu}dy^\mu dy^\nu = {{\eta}}_{\alpha\beta} \boldsymbol{\omega}^{\alpha}\boldsymbol{\omega}^{\beta}. \end{equation} This shows that the frame $\{\boldsymbol{\omega}^\mu\}_{\mu=0}^3$ is indeed orthonormal. With respect to this frame the electromagnetic Sommerfeld potential \refeq{def:AQI} becomes $\mathbf{A} = \tilde{A}_\mu \boldsymbol{\omega}^\mu$, with \begin{equation}\label{def:Atilde} \tilde{A}_0 = -Q\frac{r}{|\rho|\varpi} - \left(Q-{I\pi a}\right)\frac{a^2r\sin^2\theta}{\varpi |\rho|^3},\quad \tilde{A}_1 = 0, \quad \tilde{A}_2 = - \left(Q-I\pi a\right)\frac{ar\sin\theta}{|\rho|^3},\quad \tilde{A}_3 = 0. \end{equation} \begin{rem}\textit{ We observe that for $Q =I\pi a$, all but one of the quantities $\tilde{A}_\mu$ vanish, and the non-vanishing one, $\tilde{A}_0$, reduces to $-{Qr}/{|\rho|\varpi}$.} \end{rem} \begin{rem}\textit{ Clearly, the Cartan co-frame (\ref{CcoF}) is not invariant under the replacement $a\to-a$; it is adapted to (\ref{def:AKN}). If we replace $a\to-a$ in (\ref{def:AKN}), we also need to replace $a\to-a$ in (\ref{CcoF}); in the same vein, replacing $a\to-a$ in (\ref{CcoF}) corresponds to replacing $I\to-I$ in (\ref{def:AQI}). Therefore, in the following, a change $a\to-a$ needs to be simultaneously accompanied by the change $I\to-I$.} \end{rem} Next, let the frame of vector fields $\{{\mathbf{e}}_{\mu}\}$ be the {\em dual} frame to $\{\boldsymbol{\omega}^{\mu}\}$. Thus $\{\mathbf{e}_\mu\}$ yields an orthonormal basis for the tangent space at each point in the manifold: \begin{equation} \mathbf{e}_0 = \frac{\varpi}{|\rho|} \partial^{}_t + \frac{a}{\varpi|\rho|} \partial^{}_\varphi,\quad \mathbf{e}_1 = \frac{1}{|\rho|}\partial^{}_\theta,\quad \mathbf{e}_2 = \frac{a\sin\theta}{|\rho|} \partial^{}_t + \frac{1}{|\rho|\sin\theta} \partial^{}_\varphi,\quad \mathbf{e}_3 = \frac{\varpi}{|\rho|} \partial^{}_r\;. \end{equation} Next, the anti-symmetric matrix $\big(\boldsymbol{\Omega}_{\mu\nu}\big) = \big({{\eta}}_{\mu\lambda}\boldsymbol{\Omega}^\lambda_\nu\big)$ is computed to be \begin{equation} (\boldsymbol{\Omega}_{\mu\nu}) = \left(\begin{array}{cccc} 0&-C\boldsymbol{\omega}^0 - D\boldsymbol{\omega}^2 & D\boldsymbol{\omega}^1 - B\boldsymbol{\omega}^3 & -A\boldsymbol{\omega}^0- B \boldsymbol{\omega}^2\\ & 0 & D\boldsymbol{\omega}^0 + F\boldsymbol{\omega}^2 &-E \boldsymbol{\omega}^1 - C\boldsymbol{\omega}^3 \\ &\mbox{(anti-sym)}& 0 & -B \boldsymbol{\omega}^0 - E\boldsymbol{\omega}^2 \\ & & & 0 \end{array}\right), \end{equation} with \begin{equation} A := \frac{a^2 r \sin^2\theta}{\varpi |\rho|^3},\, B := \frac{a r \sin\theta}{|\rho|^3},\, C := \frac{a^2 \sin\theta\cos\theta}{|\rho|^3},\, D := \frac{a\cos\theta\varpi}{|\rho|^3},\, E := \frac{r\varpi}{|\rho|^3},\, F := \frac{\varpi^2\cos\theta}{|\rho|^3\sin\theta}. \end{equation} With respect to this frame on a zero-$G$ Kerr spacetime, and picking the {\em Weyl} representation\footnote{Here and throughout this paper we use the {\em Weyl} (spinor) representation for the gamma matrices, see \cite{ThallerBOOK} for details.} for the Dirac matrices $\gamma^\mu$, the covariant derivative part of the Dirac operator (\ref{eq:DirEqA}) can be expressed with the help of the operator \begin{equation} \mathfrak{O} := \tilde{\gamma}^\mu\nabla_\mu = \left(\begin{array}{cc} 0 & \mathfrak{l}'+\mathfrak{m}'\\ \mathfrak{l}+\mathfrak{m} &0 \end{array}\right), \end{equation} where \begin{equation} \mathfrak{l} := \frac{1}{|\rho|} \left(\begin{array}{cc} D_+ & L_- \\ L_+ & D_-\end{array}\right) \end{equation} and \begin{equation} \mathfrak{l}' := \frac{1}{|\rho|} \left(\begin{array}{cc} D_- & -L_- \\ -L_+ & D_+\end{array}\right), \end{equation} with \begin{equation}\label{eq:DpmLpm} D_\pm := \pm \varpi \partial^{}_r + \left( \varpi \partial^{}_t + \frac{a}{\varpi} \partial^{}_\varphi\right), \qquad L_\pm :=\partial^{}_\theta \pm i \left(a \sin\theta\,\partial^{}_t+\csc\theta \partial^{}_\varphi\right), \end{equation} while \begin{equation} \begin{aligned} \mathfrak{m} &:= \frac{1}{2}\bigl[ (-2C+F+iB)\sigma^{}_1+(-A+2E+iD)\sigma^{}_3\bigr] \\ &\ = \frac{1}{2|\rho|} \left(\begin{array}{cc} \frac{r}{\varpi}+ \frac{\varpi}{\bar{\rho}} &\cot\theta + \frac{ia\sin\theta}{\bar{\rho}}\\ \cot\theta + \frac{ia\sin\theta}{\bar{\rho}} & -\frac{r}{\varpi} - \frac{\varpi}{\bar{\rho}}\end{array}\right) \end{aligned} \end{equation} and \begin{equation} \mathfrak{m}': = \frac{1}{2}\bigl[ (2C-F+iB)\sigma^{}_1+(A-2E+iD)\sigma^{}_3\bigr] = -\mathfrak{m}^*, \end{equation} where the $\sigma^{}_k$ are Pauli matrices: \begin{equation} \sigma^{}_1 = \left(\begin{array}{cc} 0 & 1\\ 1 & 0\end{array}\right),\quad \sigma^{}_2 = \left(\begin{array}{cc} 0 & -i\\ i &\ 0\end{array}\right),\quad \sigma^{}_3 = \left(\begin{array}{cc} 1 & \ 0\\ 0 & -1\end{array}\right). \end{equation} We note that the principal part of $|\rho|\mathfrak{O}$ has an additive separation property: \begin{equation}\label{eq:DDprincipal} \begin{aligned} |\rho|\left(\begin{array}{cc} 0 & \mathfrak{l}'\\ \mathfrak{l} & 0\end{array}\right) = \left[ \gamma^3 \varpi \partial^{}_r + \gamma^0\left((\varpi \partial^{}_t + \frac{a}{\varpi} \partial^{}_\varphi\right)\right] + \Bigl[ \gamma^1 \partial^{}_\theta + \gamma^2(a\sin\theta \partial^{}_t + \csc\theta\,\partial^{}_\varphi)\Bigr], \end{aligned} \end{equation} where the coefficients of the two square-bracketed operators are functions of only $r$, respectively only $\theta$. Moreover, it is possible to transform away the lower order term in $\mathfrak{O}$, so that exact separation can be achieved for $|\rho|\mathfrak{O}$. Namely, let \begin{equation} \chi(r,\theta) := \frac{1}{2} \log( \varpi \bar{\rho}\sin\theta). \end{equation} It is easy to see that \begin{equation} \mathfrak{m} = \mathfrak{l}\chi,\qquad \mathfrak{m}' = \mathfrak{l}'\bar{\chi}. \end{equation} Let us therefore define the diagonal matrix \begin{equation}\label{def:D} \mathfrak{D} := \mbox{diag}( e^{-\chi},e^{-\chi}, e^{-\bar{\chi}}, e^{-\bar{\chi}}) \end{equation} and a new bispinor $\hat{\Psi}$ related to the original $\Psi$ by \begin{equation} \Psi = \mathfrak{D} \hat{\Psi}. \end{equation} Denoting the upper and lower components of a bispinor $\Psi$ by $\psi_1$ and $\psi_2$ respectively, it then follows that \begin{equation} (\mathfrak{l} + \mathfrak{m})\psi_1 = (\mathfrak{l} + \mathfrak{m})(e^{-\chi}\hat{\psi}_1) = e^{-\chi} \left[ \mathfrak{l} - \mathfrak{l}\chi + \mathfrak{m}\right]\hat{\psi}_1 = e^{-\chi} \mathfrak{l}\hat{\psi}_1, \end{equation} and similarly \begin{equation} (\mathfrak{l}'+ \mathfrak{m}')\psi_2 = e^{-\bar{\chi}} \mathfrak{l}'\hat{\psi}_2. \end{equation} We now put it all together. We set \begin{equation} \mathfrak{R} := \mbox{diag}(\rho,\rho,\bar{\rho},\bar{\rho}) \end{equation} and note that $|\rho|\mathfrak{D}^{-*}\mathfrak{D} = \mathfrak{R}$ while $\mathfrak{D}^{-*}\gamma^\mu\mathfrak{D} = \gamma^\mu$. Thus, setting $\Psi = \mathfrak{D} \hat{\Psi}$ in \refeq{eq:DirEqA} and left-multiplying the equation by the diagonal matrix $\mathfrak{D}' := |\rho|\mathfrak{D}^{-*}$ we conclude that $\hat{\Psi}$ solves a new Dirac equation \begin{equation}\label{eq:newDir} \left(|\rho|\gamma^\mu (\mathbf{e}_\mu + ie\tilde{A}_\mu) + im\mathfrak{R}\right) \hat{\Psi} = 0. \end{equation} Finally, let us compute the Hamiltonian form of \refeq{eq:newDir}. Let matrices $M^\mu$ be defined by \begin{equation} |\rho|\gamma^\mu\mathbf{e}_\mu = M^\mu\partial^{}_\mu. \end{equation} Thus in particular \begin{equation} M^0 = \varpi\gamma^0 + a \sin\theta\, \gamma^2. \end{equation} We may thus rewrite \refeq{eq:newDir} as \begin{equation} M^0 \partial^{}_t\hat{\Psi} = - \left( M^k\partial^{}_k + ie|\rho|\gamma^\mu\tilde{A}_\mu + im\mathfrak{R}\right)\hat{\Psi}, \end{equation} so that, defining \begin{equation}\label{def:Hhat} \hat{H} := -i (M^0)^{-1} \left( M^k\partial^{}_k + ie|\rho|\gamma^\mu\tilde{A}_\mu + im\mathfrak{R}\right), \end{equation} we can now rewrite the Dirac equation \refeq{eq:newDir} in Hamiltonian form: \begin{equation}\label{eq:DIRACeqHAMformat} i \partial^{}_t \hat{\Psi} = \hat{H}\hat{\Psi}. \end{equation} \begin{rem}\textit{ We note that for $Q \ne I\pi a$ the quantity $|\rho|\gamma^\mu\tilde{A}_\mu$ in \refeq{def:Hhat} is a function of both $r$ and $\theta$, and unlike the other terms in the Dirac equation \refeq{eq:newDir} it does {\em not} separate into a sum of two terms each depending only on one of these variables. It follows that the Dirac equation will not be exactly separable in its four spacetime variables on zGK spacetimes decorated with a generalized Sommerfeld field having a magnetic moment different from $Qa$.} \textit{Even when $Q =I\pi a$, so that $|\rho|\gamma^\mu\tilde{A}_\mu$ reduces to $|\rho|\gamma^0\tilde{A}_0 = -({Qr}/{\varpi})\gamma^0$, which is a function of only $r$, the separation of variables Ansatz does not yield a system of ordinary differential equations which can be solved one at a time, unlike the situation for the familiar Dirac equation for the spectrum of Hydrogen in Minkowski spacetime.} \end{rem} \subsubsection{A Hilbert space for $\hat{H}$} In order to decide what is the correct inner product to use for the space of bispinor fields defined on the z$G$KN spacetime, we pause to consider the action for the original Dirac equation \refeq{eq:DirEqA}, which should be obtainable from this equation upon left-multiplying it by the {\em conjugate bispinor} $\overline{\Psi}$, defined as \begin{equation} \overline{\Psi} := \Psi^\dag \gamma^0, \end{equation} and integrating the result on the spacetime. Note that the dagger in the above formula is the usual notation for ``conjugate-transpose'', i.e. $\Psi^\dag = \Psi^{*t}$, and that $\gamma^0$ is the zero-th Dirac gamma matrix for the Minkowski space, defined by \refeq{MinkGamma}.\footnote{On the Minkowski space, the $\gamma^0$ matrix plays a double role: In addition to being one of the four Dirac gamma matrices, it is also the matrix of the Hermitian quadratic form of signature (2,2) defined on the bispinor space, which is one of the key geometric structures needed in order to define the Dirac operator on a general four-dimensional Lorentzian manifold, see \cite{CanJad98} for details. In our context, the first role is played by $\tilde{\gamma}^0$, and the second one by $\gamma^0$.} Thus, using oblate spheroidal coordinates, \begin{equation} {\mathcal S}[\Psi] = \int dt \int_{\Sigma_t} \Psi^\dag \gamma^0 \left[ \tilde{\gamma}^\mu \nabla_\mu \Psi + \dots \right] d\mu^{}_{\Sigma_t}, \end{equation} where \begin{equation}\label{vol-elem} d\mu^{}_{\Sigma_t} = |\rho|^2\sin\theta d\theta d\varphi dr \end{equation} is the volume element of $\Sigma_t$, the spacelike $t=$ constant slice of z$G$KN. It follows that the natural inner product for bispinors on $\Sigma_t$ needs to be \begin{equation} \langle \Psi,\Phi\rangle = \int_\Sigma \Psi^\dag\gamma^0\tilde{\gamma}^0 \Phi d\mu^{}_\Sigma = \int_0^{2\pi}\int_0^\pi \int_{-\infty}^\infty \Psi^\dag M \Phi |\rho|^2 \sin\theta d\theta d\varphi dr, \end{equation} with \begin{equation} M := \gamma^0 \tilde{\gamma}^0 = \gamma^0 \mathbf{e}_\nu^0 \gamma^\nu = \frac{\varpi}{|\rho|} \alpha^0 + \frac{a\sin\theta}{|\rho|} \alpha^2. \end{equation} Here, $\alpha^2$ is the second one of the three Dirac alpha matrices in the Weyl (spinor) represenation, viz. \begin{equation} \alpha^k = \gamma^0 \gamma^k = \left(\begin{array}{cc} \sigma^{}_k & \ 0\\ 0 & -\sigma^{}_k\end{array}\right),\qquad k=1,2,3; \end{equation} for notational convenience, we have also set \begin{equation} \alpha^0 = \left(\begin{array}{cc} \boldsymbol{1}_{2\times2} & 0 \\ 0 & \boldsymbol{1}_{2\times2}\end{array}\right) \end{equation} for the $4\times4$ identity matrix. Now, let $\Psi = \mathfrak{D} \hat{\Psi}$ and $\Phi = \mathfrak{D} \hat{\Phi}$, with $\mathfrak{D}$ as in \refeq{def:D}. Then we have \begin{equation} \langle \Psi,\Phi \rangle = \int_{-\infty}^\infty \int_0^{2\pi} \int_0^\pi \hat{\Psi}^\dag \hat{M} \hat{\Phi} d\theta d\varphi dr, \end{equation} where \begin{equation} \hat{M} := \alpha^0 + \frac{a\sin\theta}{\varpi} \alpha^2. \end{equation} The eigenvalues of $\hat{M}$ are $\lambda_\pm = 1 \pm \frac{a\sin\theta}{\varpi}$, both of which are positive everywhere on this space with Zipoy topology. (Note that $\lambda_- \to 0$ on the ring, which is not part of the space time but at its boundary.) We may thus take the above as the definition of a positive definite inner product given by the matrix $\hat{M}$ for bispinors defined on the rectangular cylinder ${\mathcal{Z}} :={\mathbb R}\times [0,\pi]\times [0,2\pi]$ (which is the $t=const.$ section of $\mathcal{C}$) with its natural measure: \begin{equation}\label{def:innerPROD} \langle \hat{\Psi},\hat{\Phi}\rangle_{\hat{M}} := \int_{{\mathcal{Z}}} \hat{\Psi}^\dag\hat{M} \hat{\Phi} d\theta d\varphi dr. \end{equation} An alternative way of arriving at this inner product is to define the conserved Dirac current \begin{equation}\label{dir-curr} j^\mu = \overline{\Psi}\tilde{\gamma}^\mu \Psi = \Psi^\dag \gamma^0 \tilde{\gamma}^\mu \Psi, \end{equation} and consider the integral of its time component $j^0$ on the Cauchy hypersurface $\Sigma_t$ with its induced measure \refeq{vol-elem}: \begin{equation} \int_{\Sigma_t} j^0 d\mu_\Sigma = \int_{-\infty}^\infty \int_0^{2\pi} \int_0^\pi \Psi^\dag \gamma^0\tilde{\gamma}^0 \Psi |\rho|^2 \sin\theta d\theta d\varphi dr = \int_{-\infty}^\infty \int_0^{2\pi} \int_0^\pi \hat{\Psi}^\dag \hat{M}\hat{\Psi} d\theta d\varphi dr. \end{equation} The corresponding Hilbert space is denoted by ${\sf H}$, thus \begin{equation} {\sf H} := \left\{ \hat\Psi:{\mathcal{Z}} \to {\mathbb C}^4\ | \ \|\hat\Psi\|_{\hat{M}}^2 := \langle \hat\Psi,\hat\Psi\rangle_{\hat{M}} < \infty \right\}. \end{equation} Note that ${\sf H}$ is \emph{not equivalent} to standard $L^2({\mathcal{Z}})$ whose inner product has the identity matrix in place of $\hat{M}$. After these preparations we are now ready to state our main results. \subsection{Statement of the Main Theorems} Our results about the symmetry of the spectrum are valid for the Dirac Hamiltonian on a static spacelike slice of the zero-$G$ Kerr spacetime decorated with Sommerfeld fields of arbitrary charge $Q$ and current $I$. The essential self-adjointness, and location of essential and point spectra, are stated only for the Dirac Hamiltonian on a static spacelike slice of the z$G$KN spacetime; however, we conjecture that these results also hold for the more general Hamiltonian as long as the coupling constant $(Q -I\pi a)e$ is sufficiently small. In the ensuing four sections we will prove the following Theorems about $\hat{H}$. \subsubsection{Symmetry of the spectrum of the Dirac Hamiltonians} We shall find an operator which anti-commutes with any self-adjoint extension of the formal Dirac operator $\hat{H}$ on $\sf H$, with the help of which we prove: \begin{thm}\label{thm:sym} Let any self-adjoint extension of the formal Dirac operator $\hat{H}$ on $\sf H$ be denoted by the same letter. Suppose $E\in\mathrm{spec}\,\hat{H}$. Then $-E\in\mathrm{spec}\,\hat{H}$. \end{thm} Note that the above result holds for \emph{any} self-adjoint extension of $\hat{H}$, whatever $Q$ and $I$ are. \subsubsection{Essential self-adjointness of the Dirac Hamiltonian on z$G$KN} Let ${\mathcal{Z}}^*$ denote ${\mathcal{Z}}$ with the ring singularity removed. By adapting an argument of Winklmeier--Yamada \cite{WINKLMEIERc}, we shall prove: \begin{thm}\label{thm:esa} For $Q =I\pi a$, i.e. for z$\,G$KN, the operator $\hat{H}$ with domain $C^\infty_c({\mathcal{Z}}^*,{\mathbb C}^4)$ is e.s.a. in~$\sf H$. \end{thm} \subsubsection{The continuous spectrum of the Dirac Hamiltonians on z$G$KN} By adapting an argument of Weidmann \cite{Wei82}, we shall prove: \begin{thm}\label{thm:essspec} For $Q=I\pi a$ the continuous spectrum of $\hat{H}$ on $\sf H$ is ${\mathbb R}\setminus(-m,m)$. \end{thm} \subsubsection{The point spectrum of the Dirac Hamiltonian on z$G$KN} With the help of the Chandrasekhar--Page--Toop formalism to separate variables, and the Pr\"ufer transform, we will be able to control the point spectrum for the z$G$KN Dirac Hamiltonian: \begin{thm}\label{thm:ptspec} Suppose $Q =I\pi a$. Then, if $2m|a|<1$ and $|eQ|<\sqrt{2m|a|(1-2m|a|)}$, the point spectrum of $\hat{H}$ on $\sf H$ is nonempty and located in $(-m,m)$; the end points are not included. \end{thm} This completes the formulation of our main results. We next turn to their proofs. The proofs of our main theorems are distributed over four sections corresponding to the various aspects of the spectrum, i.e. symmetry, essential self-adjointness, continuous spectrum, and point spectrum. \section{Proof of Theorem \ref{thm:sym} (Symmetry of the energy spectrum)}\label{sec:proofofsymmetry} Suppose $E\in{\mathbb R}$ is an eigenvalue of $\hat{H}$. Then there exists $\hat\Psi\in {\sf H}$ such that \begin{equation} \hat{H} \hat\Psi = E \hat\Psi. \end{equation} Suppose one can find a bounded linear, or conjugate-linear, operator $\hat{C}:{\sf H} \to {\sf H}$ that anti-commutes with $\hat{H}$, i.e. \begin{equation} \bigl[ \hat{C}, \hat{H}\bigr]_+^{} = \hat{C} \hat{H} + \hat{H}\hat{C} = 0. \end{equation} It is then easy to see that $-E$ must also be an eigenvalue of $\hat{H}$, since \begin{equation} \hat{H} \hat{C} \hat\Psi = - \hat{C} \hat{H} \hat\Psi = -\hat{C} E \hat\Psi = -E \hat{C}\hat\Psi. \end{equation} This argument can be extended to show the symmetry of other parts of the spectrum. (See e.g. Glazman \cite{Gla65}, p. 205.) Let $\hat{K}:{\sf H} \to {\sf H}$ denote the complex conjugation opertor $\hat{K} \hat\Psi (x) = \hat\Psi^*(x)$, and let $\hat{S}:{\sf H}\to {\sf H}$ denote the operator $(\hat{S} \hat\Psi) (x) = \hat\Psi (\varsigma(x))$ where $\varsigma:\mathcal{Z} \to \mathcal{Z}$ is the sheet swapping map, \begin{equation} \varsigma(r,\theta,\varphi) = (-r,\pi-\theta,\varphi). \end{equation} We claim that the operator $\hat{C}:{\sf H} \to {\sf H}$ given (in Weyl representation) by $\hat{C} := \gamma^0 \hat{K} \hat{S}$, viz. \begin{equation} (\hat{C} \hat\Psi) (x) = \gamma^0 \hat\Psi^*(\varsigma(x)), \end{equation} anti-commutes with $\hat{H}$. Note that the double-sheetedness of the underlying space plays an essential role in the definition of this operator. \begin{rem}\emph{ The operator $\hat{C}$ should not be confused with the operator $\tilde{C}$ given in Weyl representation by $$ \tilde{C} := i\gamma^2 \hat{K}. $$ One easily checks that if $\hat\Psi$ solves $i\hbar\partial_t\hat\Psi = (\hat{H}_0 + e\mathcal{A})\hat\Psi$, then $\tilde{C}\hat\Psi$ solves $i\hbar\partial_t(\tilde{C}\hat\Psi) = (\hat{H}_0 - e\mathcal{A})(\tilde{C}\hat\Psi)$. In particular, if $\hat\Psi$ is an eigen-bi-spinor of $\hat{H}_0 + e\mathcal{A}$ with eigenvalue $E$, then $\tilde{C}\hat\Psi$ is an eigen-bi-spinor of $\hat{H}_0 - e\mathcal{A}$ with eigenvalue $-E$ (note that the two Hamiltonians here are different!). For this reason $\tilde{C}$ is called the {\em charge conjugation} operator. } \end{rem} To prove the claim, first note that $\gamma^0 = \beta$ anti-commutes with all three $\alpha^k$ matrices. Recall that \begin{equation}\label{def:hamhat} \hat{H}(x) = \hat{M}^{-1}\mathfrak{H} \end{equation} and \begin{equation}\label{def:hamfrak} \mathfrak{H} := -i\alpha^3 \partial^{}_r + \frac{1}{\varpi} \Bigl( - i \alpha^1 \partial^{}_\theta - i\alpha^2 \csc\theta\,\partial^{}_\varphi\Bigr) - \frac{ia}{\varpi^2} \alpha^0 \partial^{}_\varphi + \frac{m}{\varpi} \gamma^0 \mathfrak{R} + \frac{e|\rho|}{\varpi} \left(\tilde{A}^0(x) \alpha^0 + \tilde{A}^2(x)\alpha^2\right). \end{equation} Now \begin{equation}\label{minv} \hat{M}^{-1} = \frac{\varpi^2}{|\rho|^2} \left(\alpha^0 - \frac{a\sin\theta}{\varpi} \alpha^2\right). \end{equation} Thus, keeping in mind that $\overline{\alpha^2} = -\alpha^2$, we find that \begin{equation} \hat{C} \hat{M}^{-1} = \gamma^0 \overline{\hat{M}^{-1}\circ\varsigma} \hat{K}\hat{S}= \frac{\varpi^2}{|\rho|^2} \gamma^0\left(\alpha^0 +\frac{a\sin\theta}{\varpi}\alpha^2\right) \hat{K}\hat{S}= \frac{\varpi^2}{|\rho|^2}\left(\alpha^0 - \frac{a\sin\theta}{\varpi} \alpha^2\right)\gamma^0\hat{K}\hat{S}= \hat{M}^{-1}\hat{C}. \end{equation} So we only need to check that $\hat{C}$ anti-commutes with $\mathfrak{H}$. It is enough to check that each term in $\mathfrak{H}$ goes through an odd number of sign changes (either one or three) as the three operators $\hat{K}$, $\hat{S}$, and multiplication by $\gamma^0$, filter through that term. Recalling that the potential $\mathbf{A}$ is anti-symmetric with respect to sheet swap: $\tilde{A}_\mu\circ\varsigma = - \tilde{A}_\mu$, this becomes obvious for most terms in $\mathfrak{H}$. Only the term involving $\mathfrak{R}$ requires some care. We first check that $\mathfrak{R}\circ\varsigma = -\mathfrak{R}$ and that $\gamma^0 \overline{\mathfrak{R}} = \mathfrak{R} \gamma^0$. Then \begin{equation} \hat{C} \gamma^0 \mathfrak{R} = \gamma^0 \gamma^0 \overline{\mathfrak{R}\circ\varsigma} \hat{K} \hat{S} = - \gamma^0 \gamma^0 \overline{\mathfrak{R}} \hat{K} \hat{S}= - \gamma^0 \mathfrak{R} \gamma^0 \hat{K} \hat{S} = - \gamma^0\mathfrak{R}\hat{C}, \end{equation} establishing the anti-commutation property. The proof of Theorem \ref{thm:sym} is complete. Before moving on to the proof of the next theorem on the list, we pause briefly to recall our earlier discussion that showed that the physics does not change if in the z$G$KN electromagnetic spacetime solution one changes $a\to-a$, respectively changes $I\to-I$ or $a\to-a$ in its generalization involving $Qa\to I\pi a^2$. This suggests that the spectrum of our Dirac Hamiltonian must be invariant under these transformations. However, recall that our co-frame is adapted to the electromagnetic fields written as in (\ref{def:AKN}), respectively (\ref{def:AQI}), so that a sign change $a\to-a$ needs to be accompanied by a sign change $I\to-I$ in order for the physics (here: the spectrum of the Hamiltonian) to remain unchanged. We now use a variant of the strategy of proof of Theorem \ref{thm:sym} to prove exactly this. Thus we write the Hamiltonian defined by \refeq{def:hamhat} as $\hat{H} = \hat{H}_{a,I}$ to emphasize the dependence on the two parameters $a$ and $I$. \begin{prop}\label{prop:apos} There exists an involutive isometry $C: {\sf H} \to {\sf H}$ such that \begin{equation}\label{intertwine} C \hat{H}_{a,I} = \hat{H}_{-a,-I} C. \end{equation} Thus, the spectral properties of the two Hamiltonians are identical. \end{prop} \begin{proof} Let $T: {\sf H} \to {\sf H}$ be defined by $$ T \hat{\Psi} (r,\theta,\varphi) = \hat{\Psi}(r,\pi - \theta,-\varphi), $$ and let $\vec{S}$ denote the ``classical spin operator,'' in Weyl representation given by: $$ S_k := \left( \begin{array}{cc} \sigma_k & \\ & \sigma_k \end{array}\right),\qquad k=1,2,3. $$ It can then be readily checked from \refeq{def:hamhat}, \refeq{def:hamfrak}, \refeq{minv}, and \refeq{def:Atilde} that $$ C := S_3 T $$ will do the job. \end{proof} \section{Proof of Theorem \ref{thm:esa} (Essential self-adjointness ($Q=I\pi a$))} We now show that the Dirac Hamiltonian $\hat{H}$ is essentially self-adjoint on $\sf H$; recall that $\sf H$ is equipped with the inner product\footnote{As pointed out by the anonymous referee, an alternate proof of essential self-adjointness might be possible following the strategy of Chernoff \cite{Che73}, who proved essential self-adjointness of the Dirac operator on certain {\em complete} spacetimes. Such a proof would be highly welcome, indeed, for it would avoid a partial wave decomposition and be more direct. However, due to the presence of the naked ring singularity in the z$G$KN spacetime, the underlying manifold is not complete and thus it is far from obvious how to generalize Chernoff's result.} \refeq{def:innerPROD}. We observe that $M^0 = \varpi \gamma^0 \hat{M}$, so we may rewrite \refeq{def:Hhat} as \begin{equation} \hat{H} = \hat{M}^{-1}\gamma^0\left(\frac{-i}{\varpi} M^k\partial^{}_k + e\frac{|\rho|}{\varpi}\gamma^\mu\tilde{A}_\mu + \frac{m}{\varpi} \mathfrak{R}\right) = \hat{M}^{-1} \mathfrak{H}, \end{equation} where \begin{equation} \mathfrak{H} := \mathfrak{M} + \frac{1}{\varpi} \mathfrak{N} + \mathfrak{P} +\mathfrak{Q}, \end{equation} with \begin{eqnarray} \mathfrak{M} & := & -i \alpha^3 \partial^{}_r \\ \mathfrak{N} & :=& -i \alpha^1 \partial^{}_\theta -i \alpha^2\csc\theta \partial^{}_\varphi\\ \mathfrak{P} & := & -i \frac{a}{\varpi^2}\alpha^0 \partial^{}_\varphi + \frac{m}{\varpi} \gamma^0\mathfrak{R} \\ \mathfrak{Q} & := & e\frac{|\rho|}{\varpi}\gamma^0\gamma^\mu\tilde{A}_\mu\,. \end{eqnarray} Thus, \begin{equation} \langle \hat{\Psi} , \hat{H} \hat{\Phi}\rangle_{\hat{M}} = \int_{{\mathcal{Z}}} \hat{\Psi} \mathfrak{H} \hat{ \Phi} d\theta d\varphi dr. \end{equation} Evidently, $\mathfrak{H}$ is Hermitian symmetric on the Hilbert space $L^2({\mathcal{Z}};{\mathbb C}^4)$ with its natural inner product \begin{equation} \label{def:IP} (\hat\Phi,\hat\Psi) = \int_{{\mathcal{Z}}} \hat\Phi^\dag\hat\Psi d\theta d\varphi dr. \end{equation} It is furthermore easy to see that $\hat{H}$ is e.s.a. on ${\sf H}$ if and only if $\mathfrak{H}$ is e.s.a. on $L^2({\mathcal{Z}};{\mathbb C}^4)$. We shall prove that $\mathfrak{H}$ is e.s.a. on $L^2({\mathcal{Z}};{\mathbb C}^4)$ when $Q=I\pi a$, i.e. for a Dirac point electron in z$G$KN. \begin{thm}\label{thm:esaGOTIC} For $Q=I\pi a$ the operator $\mathfrak{H}$ with domain $C^\infty_c({\mathcal{Z}}^*,{\mathbb C}^4)$ is e.s.a. in $L^2({\mathcal{Z}},{\mathbb C}^4)$. \end{thm} \begin{proof} Let us write \begin{equation} \mathfrak{H} = \mathfrak{H}^0 + \mathfrak{Q}. \end{equation} Here, $\mathfrak{H}^0$ is the free Hamiltonian. We will first show that $\mathfrak{H}^0$ is essentially self-adjoint; this proof is an easy adaptation of the method first employed by Winklmeier and Yamada \cite{WINKLMEIERc}. We will then conclude essential self-adjointness of $\mathfrak{H}$ by using a perturbation argument. To this end, let us consider the decomposition with respect to the azimuthal angle $\varphi$ of the Hilbert space $L^2({\mathcal{Z}};{\mathbb C}^4)$ into partial wave subspaces $L^2([0,\pi]\times{\mathbb R},d\theta dr)$: \begin{equation}\label{pwd} L^2({\mathcal{Z}};{\mathbb C}^4) = \oplus_{\kappa\in{\mathbb Z}+\frac{1}{2}} \left(L^2_\kappa([0,\pi]\times{\mathbb R},d\theta dr)\right)^4 \end{equation} corresponding to the expansion of a bispinor field $\hat\Psi \in L^2({\mathcal{Z}};{\mathbb C}^4)$ given by \begin{equation}\label{decomp} \hat\Psi (r,\theta,\varphi) = \sum_{\kappa\in{\mathbb Z}+\frac{1}{2}} e^{i\kappa\varphi}\hat\Psi_\kappa(r,\theta). \end{equation} For a discussion of why $\kappa$ needs to be a half-integer, see \cite{FinsterETalDperDNE,FinsterETalDperDNEerr}. Let $\mathfrak{H}^0_\kappa := \left.\mathfrak{H}^0\right|_{L^2_\kappa}$. Then $\mathfrak{H}^0_\kappa = \mathfrak{S}_\kappa + \varpi^{-1}\mathfrak{T}_\kappa +\mathfrak{B}_\kappa$, with \begin{equation} \mathfrak{S}_\kappa = -i \alpha^3 \partial^{}_r , \qquad \mathfrak{T}_\kappa =-i \alpha^1 \partial^{}_\theta + \alpha^2\kappa \csc\theta = \left(\begin{array}{cc} \mathfrak{t}_\kappa& 0\\ 0 & -\mathfrak{t}_\kappa\end{array}\right), \qquad \mathfrak{B}_\kappa = \frac{a\kappa}{\varpi^2}\alpha^0 + \frac{m}{\varpi} \gamma^0\mathfrak{R}. \end{equation} We note that $\mathfrak{B}_\kappa$ is a symmetric {\em bounded} multiplication operator on $L^2_\kappa$; in fact, \begin{equation} \|\mathfrak{B}_\kappa \|_{L^\infty} \leq |{\kappa}/{a}| +m, \end{equation} so that the task of showing e.s.a.-ness of $\mathfrak{H}^0_\kappa$ reduces to showing e.s.a.-ness of $\mathfrak{H}'_\kappa := \mathfrak{S}_\kappa+\varpi^{-1}\mathfrak{T}_\kappa$. Now $\mathfrak{H}'_\kappa$ is block-diagonal: \begin{equation}\label{fHprimeDecomp} \mathfrak{H}'_\kappa = \left(\begin{array}{cc}\mathfrak{h}'_\kappa & 0 \\ 0 & -\mathfrak{h}'_\kappa \end{array}\right),\qquad \mathfrak{h}'_\kappa := -i\sigma_3 \partial_r + \varpi^{-1}\mathfrak{t}_\kappa,\qquad \mathfrak{t}_\kappa := -i \sigma_1 \partial_\theta + \sigma_2\kappa \csc\theta. \end{equation} Thus it is enough to show $\mathfrak{h}'_\kappa$ is e.s.a. We do so by showing that $\ker(\mathfrak{h}'_\kappa \pm i) = \{0\}$: Suppose $\hat\psi_\kappa \in \left(L^2_\kappa([0,\pi]\times{\mathbb R},d\theta dr)\right)^2$ satisfies \begin{equation}\label{eq:ker} \mathfrak{h}'_\kappa \hat\psi_\kappa = \pm i \hat\psi_\kappa. \end{equation} As observed in \cite{WINKLMEIERc}, it is possible to decompose \refeq{eq:ker} with respect to the eigenspaces of the operator \begin{equation} \mathfrak{a}_\kappa:= W\mathfrak{t}_\kappa W^{-1} = -i\sigma^{}_2 \partial^{}_\theta + \kappa \csc\theta \sigma^{}_1, \end{equation} where \begin{equation}\label{def:W} W := \left(\begin{array}{cc} 0 & 1\\ i & 0\end{array}\right). \end{equation} The operator $\mathfrak{a}_\kappa$ has pure point spectrum and a complete set of eigenfunctions. More precisely, one has the following result \cite{WinklmeierPHD} (here quoted from \cite{WINKLMEIERc}): \begin{thm} (Winklmeier, 2006) For all $\kappa \in {\mathbb Z} + \frac{1}{2}$ the operator $\mathfrak{a}_\kappa$ with domain $(C^\infty_c((0,\pi)))^2$ is essentially self-adjoint in $(L^2((0,\pi),d\theta))^2$. Its closure (denoted again by $\mathfrak{a}_\kappa$) is compactly invertible and its spectrum consists of simple eigenvalues only, given by \begin{equation}\label{def:lank} \lambda_n^\kappa := \mbox{sgn}(n)\left(|\kappa|-{\textstyle\frac{1}{2}} + |n|\right),\qquad n \in {\mathbb Z}^* = {\mathbb Z}\setminus\{0\}, \end{equation} with corresponding normalized eigenfunctions $\{g_n^\kappa\}_{n\in{\mathbb Z}^*}$ forming a complete orthonormal set in $(L^2((0,\pi),d\theta))^2$. Moreover, \begin{equation} \lambda_{-n}^\kappa = -\lambda_n^\kappa,\qquad g_{-n}^\kappa = -\sigma^{}_3 g_n^\kappa. \end{equation} \end{thm} We can therefore write \begin{equation}\label{winkdec} \hat\psi_\kappa(r,\theta)= \sum_{n\in{\mathbb Z}^*} \xi_n(r) g_n^\kappa(\theta), \end{equation} with functions $\xi_n\in L^2({\mathbb R},dr)$. Hence, performing a similarity transform on \refeq{eq:ker} with $W$ and projecting on the spans of $g_n^\kappa$ and $g_{-n}^\kappa$ we obtain the following system (see \cite{WINKLMEIERc} for details): \begin{equation} \left(\begin{array}{cc} \frac{\lambda_n^\kappa}{\varpi} & i \partial^{}_r \\ i\partial^{}_r & -\frac{\lambda_n^\kappa}{\varpi}\end{array}\right) \left(\begin{array}{c} \xi_n \\ \xi_{-n}\end{array}\right) = \pm i \left(\begin{array}{c} \xi_n \\ \xi_{-n}\end{array}\right) \end{equation} However the operator in the above eigenvalue problem $\mathcal{C}_\kappa = i\sigma^{}_1\partial^{}_r + \frac{\lambda_n^\kappa}{\varpi}\sigma^{}_3$ is clearly e.s.a., since $\frac{\lambda_n^\kappa}{\varpi}$ is bounded, hence $\xi_n = \xi_{-n} = 0$. This completes the proof of essential self-adjointness of $\mathfrak{H}^0$. Consider now the term $\mathfrak{Q}= e\frac{|\rho|}{\varpi}\gamma^0\gamma^\mu\tilde{A}_\mu$ coming from the electromagnetic potential. It can be rewritten as $\mathfrak{Q} = -eQ\mathfrak{V}_1 - e(Q-I\pi a)\mathfrak{V}_2$, where \begin{equation}\label{eq:fQdecomp} \mathfrak{V}_1 := \frac{r}{\varpi^2}\alpha^0,\qquad \mathfrak{V}_2 := \frac{ar\sin\theta}{\varpi|\rho|^2}\hat{M}. \end{equation} The first term, $\mathfrak{V}_1$, is clearly bounded, whereas the second one, $\mathfrak{V}_2$, blows up on the ring. However, since by hypothesis we restrict ourselves to the case $Q=I\pi a$, the $\mathfrak{V}_2$ term is absent from $\mathfrak{Q}$, and essential self-adjointness of $\mathfrak{H}$ follows easily from that of $\mathfrak{H}^0$ and the boundedness of $eQ\mathfrak{V}_1$. The proof of Theorem \ref{thm:esaGOTIC} is complete. \end{proof} For the proof the remaining statements in this paper we rely on the fact that the Dirac equation of a point electron in z$G$KN separates into four (coupled) ordinary differential equations, each of which depends on only one of the four oblate spheroidal coordinates, with the coupling being effected through shared parameters in the equations. This is carried out in the next section before we resume with proving our claims. \section{Chandrasekhar--Page--Toop separation-of-variables ($Q=I\pi a$)} When $Q=I\pi a$ the Dirac equation \refeq{eq:DIRACeqHAMformat} for the bispinor $\hat{\Psi}$ allows a clear separation also for the remaining $r$ and $\theta$ derivatives (commonly referred to in the literature as ``radial'' and ``angular'' derivatives, even though $r$ is not a radial distance and $\theta$ is not an angle, except at infinity). Thus, when $Q=I\pi a$ the Dirac equation \refeq{eq:DIRACeqHAMformat} becomes \begin{equation}\label{eq:DirSep} (\hat{R} +\hat{A}) \hat{\Psi} = 0, \end{equation} where \begin{eqnarray} \hat{R}& := & \left(\begin{array}{cccc} imr & 0 &D_-+ieQ\frac{r}{\varpi} & 0 \\ 0 & imr & 0 & D_++ieQ\frac{r}{\varpi}\\ D_++ieQ\frac{r}{\varpi} & 0 & imr & 0 \\ 0 & D_-+ieQ\frac{r}{\varpi} & 0 & imr \end{array} \right),\\ \hat{A} &:= & \left(\begin{array}{cccc} -m a \cos\theta & 0 & 0 & -L_- \\ 0 & -ma\cos\theta & -L_+ &0 \\ 0 & L_- & ma\cos\theta & 0 \\ L_+ & 0 & 0 & ma\cos\theta \end{array}\right), \end{eqnarray} where $D_\pm$ and $L_\pm$ have been given in (\ref{eq:DpmLpm}). Once a solution $\hat{\Psi}$ to \refeq{eq:DirSep} is found, the bispinor $\Psi := \mathfrak{D}\hat{\Psi}$ solves the original Dirac equation \refeq{eq:DirEqA}. \subsubsection{The Chandrasekhar Ansatz} Assume now that a solution $\hat{\Psi}$ of \refeq{eq:DirSep} is of the form \begin{equation}\label{chandra-ansatz} \hat{\Psi} = e^{-i(Et-\kappa \varphi)} \left( \begin{array}{c}R_1S_1\\ R_2 S_2\\ R_2 S_1\\ R_1 S_2 \end{array}\right), \end{equation} with $R_k$ being complex-valued functions of $r$ alone, and $S_k$ real-valued functions of $\theta$ alone. Let \begin{equation} \vec{R} := \left(\begin{array}{c} R_1\\ R_2\end{array}\right),\qquad \vec{S} := \left(\begin{array}{c} S_1\\ S_2\end{array}\right). \end{equation} Plugging the Chandrasekhar Ansatz \eqref{chandra-ansatz} into \eqref{eq:DirSep} one easily finds that there must be $\lambda\in{\mathbb C}$ such that \begin{equation}\label{eq:rad} T_{rad}\vec{R} = E\vec{R}, \end{equation} \begin{equation}\label{eq:ang} T_{ang}\vec{S} = \lambda \vec{S}, \end{equation} where \begin{eqnarray} T_{rad} & := \label{eq:Trad} & \left(\begin{array}{cc} d_- &-m\frac{r}{\varpi} - i\frac{\lambda}{\varpi} \\ -m\frac{r}{\varpi}+i\frac{\lambda}{\varpi} & -d_+ \end{array}\right) \\ T_{ang}& := \label{eq:Srad} & \left(\begin{array}{cc} -ma\cos\theta & -l_- \\ l_+ &ma\cos\theta \end{array}\right) \end{eqnarray} The operators $d_\pm$ and $l_\pm$ are now ordinary differential operators in $r$ and $\theta$ respectively, with coefficients that depend on the unknown $E$, and parameters $a$, $\kappa$, and $eQ$: \begin{eqnarray}\label{opdefs} d_\pm & := & i \frac{d}{dr} \pm \frac{-a\kappa + eQ r}{\varpi^2}\\ l_\pm & := & \frac{d}{d\theta} \mp \left( aE\sin\theta - \kappa \csc\theta\right) \end{eqnarray} The angular operator $T_{ang}$ in \refeq{eq:ang} is easily seen to be essentially self-adjoint on $(C^\infty_c((0,\pi),\sin\theta d\theta))^2$ and in fact is self-adjoint on its domain inside $(L^2((0,\pi),\sin\theta d\theta))^2$ (e.g. \cite{SufFacCos83,WINKLMEIERa}) with purely point spectrum $\lambda=\lambda_n(am,aE,\kappa)$, $n\in {\mathbb Z}\setminus 0$. Thus in particular $\lambda \in {\mathbb R}$. It then follows that the radial operator $T_{rad}$ is also essentially self-adjoint on $(C^\infty_c({\mathbb R}, dr))^2$ and in fact self-adjoint on its domain inside $(L^2({\mathbb R},dr))^2.$ Suppose $\vec{R} = (R_1,R_2)^T \in (L^2({\mathbb R}))^2$ is a nontrivial solution to $T_{rad} \vec{R} = E\vec{R}$, with $E\in {\mathbb R}$. Then \begin{eqnarray*} \frac{dR_1}{dr} - i\left(E - \frac{ a \kappa-eQr}{\varpi^2}\right)R_1 +\frac{1}{\varpi} (imr - \lambda) R_2 & = & 0\\ -\frac{d R_2}{dr} -i \left(E- \frac{ a \kappa-eQr}{\varpi^2}\right) R_2 +\frac{1}{\varpi} (imr + \lambda)R_1 & = & 0. \end{eqnarray*} Multiply the first equation by $\bar{R}_1$ and the second equation by $\bar{R}_2$, add them and take the real part, to obtain \begin{equation} \frac{d}{dr} \left(|R_1|^2 - |R_2|^2\right) = 0. \end{equation} Thus the difference of the moduli squared of $R_1$ and $R_2$ is constant, hence zero since they need to be integrable at infinity. I.e., \begin{equation} |R_1| = |R_2| := R. \end{equation} Let $R_j = R e^{i\Phi_j}$ for $j=1,2$. Multiply the first equation by $\bar{R}_2$, multiply the complex conjugate of the second equation by $R_1$, and add them to obtain \begin{equation} \frac{d}{dr} \left(\frac{R_1}{ \bar{R}_2}\right) = 0. \end{equation} Thus the ratio $R_1 / \bar{R}_2$, and hence the sum of the arguments $\Phi_1+ \Phi_2$ must be a constant, say $\delta$. Thus $R_1 = \bar{R}_2e^{i\delta}$. Since multiplication by a constant phase factor is a gauge transformation for Dirac bispinors, we can replace $\hat{\Psi}$ with $\hat{\Psi}' = e^{-i\delta/2}\hat{\Psi}$ without changing anything. The spinor thus obtained has the same form as \refeq{chandra-ansatz}, now with $R'_1 = \bar{R}'_2$. Thus without loss of generality we can assume $\delta = 0$ and $R_1 = \bar{R}_2$. This motivates us to set \begin{equation} R_1 =\frac{1}{\sqrt{2}}( v-iu),\qquad R_2 =\frac{1}{\sqrt{2}}( v + iu) \end{equation} for real funcions $u$ and $v$. Consider the unitary matrix \begin{equation} U := \frac{1}{\sqrt{2}}\left(\begin{array}{cc} -i & 1 \\ \ i & 1 \end{array}\right). \end{equation} A change of basis using $U$ brings the radial system \refeq{eq:rad} into the following standard (Hamiltonian) form \begin{equation}\label{eq:hamil} (H_{rad} -E)\left(\begin{array}{c} u \\ v \end{array}\right) = \left(\begin{array}{c}0 \\ 0 \end{array}\right), \end{equation} where \begin{equation}\label{eq:Hrad} H_{rad} := \left(\begin{array}{cc} m \frac{r}{\varpi} + \frac{\gamma r+a\kappa}{\varpi^2} & -\partial^{}_r + \frac{\lambda}{\varpi} \\[20pt] \partial^{}_r +\frac{\lambda}{\varpi} & -m\frac{r}{\varpi} + \frac{\gamma r+a\kappa}{\varpi^2} \end{array}\right), \end{equation} (cf. \cite{ThallerBOOK}, eq (7.105)) with \begin{equation} \gamma := -eQ <0. \end{equation} \section{Proof of Theorem \ref{thm:essspec} (Continuous spectrum of $\hat{H}$ on z$G$KN)}\label{sec:contspec} Following Weidmann \cite{Wei82} we now prove the theorem about the continuous spectrum of $\hat{H}$. Recall the partial wave decomposition \eqref{pwd}. Let $\hat{H}_\kappa$ denote the restriction of $\hat{H}$ to $L^2_\kappa$. The Chandrasekhar separation \refeq{chandra-ansatz} and equation \refeq{eq:rad} yield that the spectrum of $\hat{H}_\kappa$ coincides with that of $T_{rad}$, which coincides with that of $H_{rad}$ since these last two are unitarily equivalent. Furthermore, the spectrum of $\hat{H}$ equals the union of the spectra of $\hat{H}_\kappa$. Thus in order to prove the claim about the essential spectrum, it suffices to show that it holds for $H_{rad}$ regardless of the values of $\kappa$ and $\lambda$. Since $H_{rad}$ is a radial Dirac operator, one can then use results that are particular to one dimension. One such result is due to Weidmann \cite{Wei82}: \begin{thm*} Let $P$ and $J$ be matrices such that $H_{rad} = J\partial_r + P$. Suppose $P$ can be written as $P_1+P_2$ in such a way that each component of $P_1$ is integrable in $[R,\infty)$ for some $R>0$, $P_2$ is of bounded variation on $[R,\infty)$ and $$ \lim_{r\to \infty} P(r) = \left(\begin{array}{cc} a & 0 \\ 0 & b \end{array} \right),\qquad a> b. $$ Then {\em each} self-adjoint extension of $h$ has a purely absolutely continuous spectrum in $(-\infty,b]\cup[a,\infty)$. \end{thm*} Using this result, our claim follows by noticing that the hypotheses on $P$ are satisfied, and $$ \lim_{r\to \infty} P(r) = \left(\begin{array}{cc} m & 0 \\ 0 & -m \end{array}\right). $$ Proof of Theorem \ref{thm:essspec} is complete. \section{Proof of Theorem \ref{thm:ptspec} (Point spectrum of $\hat{H}$ on z$G$KN)} By the remarks at the beginning of Section~\ref{sec:contspec}, we are interested in the eigenvalues $E$ and square-integrable eigenfunctions in $L^2({\mathbb R},dr)^2$ of the operator $H_\mathrm{rad}$. One complication is that in our case the radial Hamiltonian $H_{rad}$ depends on the unknown eigenvalues $\lambda$ of the angular operator $T_{ang}$ in \refeq{eq:ang}, which in turn depend on the energy $E$. Since the angular operator is the same as the one on Kerr and Kerr-Newman spacetime studied in \cite{SufFacCos83, WINKLMEIERa}, and since it is known that for a given value of $E$ there is a largest negative eigenvalue $\lambda = \Lambda(E)$, our strategy is to show the existence, for a given value of $\lambda<0$, of a smallest positive eigenvalue $E = {\mathcal E}(\lambda)$ for $H_{rad}$, and then set up an iteration that converges to a pair $(E,\lambda)$ for which the radial \refeq{eq:rad} and the angular \refeq{eq:ang} equations jointly have $L^2$ solutions, thereby establishing the existence of a ``positive-energy eigenstate'' for the full Dirac Hamiltonian; note that by the symmetry of the spectrum there also exists a ``negative-energy eigenstate.'' \subsection{The Pr\"ufer transform} Consider the equations \refeq{eq:hamil} and \refeq{eq:ang} for unknowns $(u,v)$ and $(S_1,S_2)$. Let us define new unknowns $(R,\Omega)$ and $(S,\Theta)$ via the Pr\"ufer transform \cite{Pru26} \begin{equation}\label{eq:prufer} u =\sqrt{2} R \cos\frac{\Omega}{2},\quad v = \sqrt{2} R \sin\frac{\Omega}{2},\quad S_1 = S \cos\frac{\Theta}{2},\quad S_2 = S \sin\frac{\Theta}{2}. \end{equation} Thus \begin{equation} R =\frac{1}{2}\sqrt{u^2+v^2},\quad\Omega = 2\tan^{-1}\frac{v}{u},\quad S = \sqrt{S_1^2+S_2^2},\quad \Theta = 2\tan^{-1}\frac{S_2}{S_1}. \end{equation} As a result, $R_1 = -iRe^{i\Omega/2}$ and $R_2 = iRe^{-i\Omega/2}$. Hence $\hat{\Psi}$ can be re-expressed in terms of the Pr\"ufer variables, thus \begin{equation}\label{ontology} \hat{\Psi}(t,r,\theta,\varphi) = R(r)S(\theta)e^{-i(Et-\kappa \varphi)} \left(\begin{array}{l} -i\cos(\Theta(\theta)/2)e^{+i\Omega(r)/2}\\ +i\sin(\Theta(\theta)/2) e^{-i\Omega(r)/2}\\ +i\cos(\Theta(\theta)/2)e^{-i\Omega(r)/2}\\ -i\sin(\Theta(\theta)/2)e^{+i\Omega(r)/2}\end{array}\right), \end{equation} and we obtain the following equations for the new unknowns \begin{eqnarray} \frac{d}{dr}\Omega &=& 2 \frac{mr}{\varpi} \cos\Omega + 2\frac{\lambda}{\varpi} \sin\Omega +2\frac{a\kappa + \gamma r}{\varpi^2} - 2E ,\label{eq:Om}\\ \frac{d}{dr} \ln R &=& \frac{mr}{\varpi}\sin\Omega - \frac{\lambda}{\varpi} \cos\Omega .\label{eq:R} \end{eqnarray} Similarly, \begin{eqnarray} \frac{d}{d\theta}\Theta &=& -2ma\cos\theta\cos\Theta + 2\left(aE \sin\theta - \frac{\kappa}{\sin\theta}\right)\sin\Theta + 2\lambda,\label{eq:Theta}\\ \frac{d}{d\theta} \ln S &=& -ma \cos\theta\sin\Theta - \left(aE\sin\theta - \frac{\kappa}{\sin\theta}\right)\cos\Theta. \label{eq:S} \end{eqnarray} To simplify the analysis of these systems and reduce the number of parameters involved, we will henceforth set $m=1$. Note that this is always possible by defining the constants $a'=ma$, $E'=E/m$, and a change of variable $r'=mr$. \begin{rem} \textit{Equations \eqref{eq:Om}--\eqref{eq:S} have certain symmetries that are connected with the action of operators $\hat{C}$ and $\tilde{C}$ introduced in Section~\ref{sec:proofofsymmetry}: It is easy to see that these four equations are preserved under each of the following two transformations: \begin{equation}\label{trans1} r \to -r,\quad \theta \to \pi - \theta,\quad \lambda \to -\lambda, \quad \kappa \to -\kappa, \quad E \to -E, \end{equation} and \begin{equation}\label{trans2} \Omega \to \pi - \Omega,\quad \Theta \to \pi -\Theta, \quad \lambda \to -\lambda, \quad \kappa \to -\kappa, \quad E \to -E, \quad \gamma \to -\gamma. \end{equation} The map \refeq{trans1} corresponds to the action of $\hat{C}$ and the map \refeq{trans2} to the action of $\tilde{C}$. Note that the latter does not preserve the Hamiltonian, since the sign of $\gamma = -eQ$ is changed. } \end{rem} \subsection{The realm of $L^2$ solutions} We note that in both of the above systems (\ref{eq:Om},\ref{eq:R}) and (\ref{eq:Theta},\ref{eq:S}), when a solution to the first equation is known, the second equation in the system can be solved by quadrature. Moreover, the requirement that $R$ and $S$ be $L^2$ functions of their argument determines what boundary values the solutions to the $\Omega$ and $\Theta$ equations should have. More precisely, \begin{prop}\label{prop:bndryvals} Any bispinor $\hat{\Psi}$ of the form \refeq{ontology} constructed from solutions of (\ref{eq:Om}), (\ref{eq:R}), (\ref{eq:Theta}), (\ref{eq:S}), with $|\kappa|\geq \frac{1}{2}$, $E>0$ and $\lambda<0$, belongs to the Hilbert space ${\sf H}$ provided \begin{equation}\label{asympOm} \lim_{r\to -\infty} \Omega(r) = -\pi + \cos^{-1}(E),\qquad \lim_{r\to \infty} \Omega(r) = - \cos^{-1}(E), \end{equation} and \begin{equation}\label{asympTh} \Theta(0) = 0,\qquad \Theta(\pi) =-\pi. \end{equation} \end{prop} \begin{proof} It is straightforward to compute that for a $\hat{\Psi}$ of the form \refeq{ontology}, \begin{eqnarray*} \|\hat{\Psi}\|_{\hat{M}}^2 &=& 2\int_0^{2\pi}\int_0^\pi\int_{-\infty}^\infty R^2(r)S^2(\theta)\left( 1+ \frac{a\sin\theta}{\varpi} \sin\Theta(\theta) \sin\Omega(r)\right) dr d\theta d\varphi \\ &=& 4\pi\left[ \int_{-\infty}^\infty R^2 dr \int_0^\pi S^2 d\theta + a \int_{-\infty}^\infty R^2 \sin\Omega \frac{dr}{\varpi} \int_0^\pi S^2 \sin\Theta \sin\theta d\theta\right] \\ &\leq& 8 \pi \|R\|^2_{L^2} \|S\|_{L^2}^2, \end{eqnarray*} and thus $\hat\Psi \in {\sf H}$ provided $R \in L^2({\mathbb R},dr)$ and $S \in L^2((0,\pi),d\theta)$. Now \refeq{eq:Theta} can be written as a smooth dynamical system in the $(\theta,\Theta)$ plane by introducing a new independent variable $\tau$ such that $\frac{d\theta}{d\tau} = \sin\theta$. Then, with dot representing differentiation in $\tau$, we have, \begin{equation}\label{dynsysTh} \left\{\begin{array}{rcl} \dot{\theta} & = & \sin\theta\\ \dot{\Theta} & = & -2a\sin\theta\cos\theta\cos\Theta+2aE\sin^2\theta\sin\Theta - 2\kappa\sin\Theta + 2\lambda\sin\theta \end{array}\right. \end{equation} Identifying the line $\Theta=\pi$ with $\Theta=-\pi$, this becomes a dynamical system on a closed finite cylinder $\mathcal{C}_1=[0,\pi]\times{\mathbb S}^1$. The only equilibrium points of the flow are on the two circular boundaries: Two on the left boundary: $S_- = (0,0)$, $N_- = (0,\pi)$; two on the right: $S_+ = (\pi,-\pi)$ and $N_+ = (\pi,0)$. For $\kappa>0$, the linearization of the flow at the equilibrium points reveals that $S_-$ and $S_+$ are hyperbolic saddle points (with eigenvalues $\{1,-2\kappa\}$ and $\{-1,2\kappa\}$ respectively), while $N_-$ is a source node (with eigenvalues 1 and $2\kappa$) and $N_+$ is a sink node (with eigenvalues $-1$ and $-2\kappa$). Note that the situation with $\kappa<0$ is entirely analogous, with the critical points switching their roles. For the remainder of this section therefore, we will assume $\kappa>0$. The $\alpha$-limit set of the orbit of any point in the interior of the cylinder must necessarily be either $S_-$ or $N_-$, and likewise its $\omega$-limit set can only be either $N_+$ or $S_+$. The only possible boundary values for $\Theta(\theta)$ are therefore $0$ and $\pm\pi$ at each endpoint of the interval $0\leq \theta\leq \pi$. The boundary values (\ref{asympTh}) correspond to a heteroclinic orbit connecting the two saddles $S_-$ and $S_+$. Suppose such a saddles connection exists. Since the eigendirection corresponding to the unstable manifold of $S_-$ is $v_1 = (\kappa+\frac{1}{2},\lambda-a)^T$ and the stable manifold of $S_+$ has the same eigendirection, it follows that for the said saddles connection, we have \begin{equation} \frac{d\Theta }{d\theta}_{|_{\theta=0}} = \frac{d\Theta}{d\theta}_{|_{\theta = \pi}} = \frac{\lambda -a}{\kappa+\frac{1}{2}} =:\delta <0. \end{equation} Thus $\Theta = \delta\theta + o(1)$ as $\theta\to 0$ and $\Theta = -\pi+\delta(\theta-\pi) +o(1)$ as $\theta \to \pi$. Consider now the $S$ equation \refeq{eq:S}. By the above, \begin{equation} \frac{d}{d\theta}\ln S =\left\{\begin{array}{ll} \frac{\kappa}{\theta} +o(1) & \mbox{ as }\theta \to 0\\ \frac{\kappa}{\theta-\pi} + o(1) & \mbox{ as }\theta \to \pi \end{array}\right. \end{equation} Integrating in $\theta$ we thereby conclude that $S \sim |\theta|^\kappa$ for $\theta$ small and $S \sim |\pi-\theta|^\kappa$ for $\theta$ near $\pi$. Therefore $S$ is integrable on $(0,\pi)$ and indeed it belongs to $L^p((0,\pi),d\theta)$ for any $p\geq 1$. In an analogous manner one shows that $S\not\in L^2$ when $\Theta$ takes boundary values corresponding to either of the nodes $N_-$ or $N_+$. Consider next the $\Omega$ equation \refeq{eq:Om}. It can also be rewritten as a smooth dynamical system on a cylinder, in this case by setting $\tau := \frac{r}{a}$ as new independent variable, as well as introducing a new dependent variable \begin{equation} \xi := \tan^{-1}\frac{r}{a} = \tan^{-1}\tau \end{equation} Then, with dot again representing differentiation in $\tau$, \refeq{eq:Om} is equivalent to \begin{equation}\label{dynsysOm} \left\{\begin{array}{rcl} \dot{\xi} & = & \cos^2\xi \\ \dot{\Omega} & = & 2a\sin\xi\cos\Omega+2\lambda\cos\xi\sin\Omega + 2\gamma\sin\xi\cos\xi+2\kappa\cos^2\xi - 2aE \end{array}\right. \end{equation} Once again, identifying $\Omega=-\pi$ with $\Omega = \pi$ turns this into a smooth flow on the closed finite cylinder $\mathcal{C}_2 := [-\frac{\pi}{2},\frac{\pi}{2}]\times {\mathbb S}^1$. The only equilibrium points of the flow are on the two circular boundaries. For $E\in[0,1)$ there are two equilibria on each: $S_- = (-\frac{\pi}{2},-\pi+\cos^{-1}E)$ and $N_-=(-\frac{\pi}{2},\pi - \cos^{-1}E)$ on the left boundary, and $S_+ = (\frac{\pi}{2}, -\cos^{-1}E)$ and $N_+ = (\frac{\pi}{2},\cos^{-1}E)$ on the right boundary. $S_\mp$ are non-hyperbolic (degenerate) saddle-nodes, with eigenvalues $0$ and $\pm 2a\sqrt{1-E^2}$, while $N_-$ is a degenerate source-node and $N_+$ a degenerate sink-node.\footnote{For $E=1$ each $S,N$ pair coalesces into one degenerate equilibrium: $N^1_- = (-\frac{\pi}{2},\pm\pi)$ and $N^1_+ = (\frac{\pi}{2},0)$ with both eigenvalues being zero.} The boundary values \refeq{asympOm} correspond to a heteroclinic orbit connecting $S_-$ and $S_+$. Suppose such a saddles connection exists, and consider the $R$ equation \refeq{eq:R}. As $r \to \pm\infty$, we will then have \begin{equation} \frac{d}{dr}\ln R \sim - \mbox{sgn}(r)\sqrt{1-E^2} \end{equation} so that integrating in $r$ we will obtain \begin{equation} R(r) \sim e^{-|r|\sqrt{1-E^2}}\qquad \mbox{ as } r\to \pm\infty \end{equation} which ensures that $R$ is integrable at infinity. Since the right-hand-side of the $R$ equation is smooth in $\Omega$ and $r$, and $\Omega$ itself is smooth, it follows that $R \in L^p({\mathbb R},dr)$ for all $p\geq 1$. In an analogous manner one shows that $R\not\in L^2$ when $\Omega$ takes boundary values corresponding to either of the nodes $N_-$ or $N_+$. \end{proof} \subsection{Existence of heteroclinic orbits connecting the two saddles} {F}rom the proof of Proposition~\ref{prop:bndryvals} it is evident that in order to establish the existence of an eigenfunction for the Dirac Hamiltonian of a point electron in the z$G$KN spacetime, we need to show that there exists a pair $(E,\lambda)$ such that both dynamical systems \refeq{dynsysTh} and \refeq{dynsysOm} have a {\em saddle-saddle connecting orbit} for those values of $E$ and $\lambda$. We call this type of orbit a {\em saddles connector} for the corresponding flow. We pave the road for our proof by recalling some general facts of flow on a cylinder. \subsubsection{Flow on a finite cylinder} Let $\mathcal{C} := [x_-^{},x_+^{}]\times {\mathbb S}^1$ be a finite cylinder. We denote its universal cover by $\bar{\mathcal{C}} := [x_-^{},x_+^{}]\times{\mathbb R}$, with coordinates $(x,y)$, and fix a fundamental domain $\widetilde{\mathcal{C}} := [x_-^{},x_+^{}]\times[-\pi,\pi)$ in $\bar{\mathcal{C}}$. Consider the flow $\Phi_t$ on $\bar{\mathcal{C}}$ given by the dynamical system \begin{equation}\label{eq:flow} \left\{\begin{array}{rcl} \dot{x} & = & f(x)\\ \dot{y} & = & g(x,y) \end{array}\right. \end{equation} where the dot represents differentiation with respect to a formal ``time'' parameter $\tau$, the functions $f$ and $g$ are smooth, and $g$ is $2\pi$-periodic in $y$: $g(x,y) = g(x,y+2\pi)$. Let us moreover assume that $f$ satisfies \begin{equation} f(x_-^{}) = f(x_+^{}) = 0,\qquad f(x)>0\ \forall\ x\in(x_-^{},x_+^{}) \end{equation} while $g$ satisfies \begin{equation} g(x_-^{},y) = 0 \implies y\in\{n_-^{}, s_-^{}\},\qquad g(x_+^{},y) = 0 \implies y\in\{n_+^{},s_+^{}\} \end{equation} where $-\pi\leq s_-^{} < n_-^{} \leq\pi$ and $-\pi\leq s_+^{} < n_+^{} \leq\pi$. These assumptions imply that the following four distinct points in $\mathcal{C}$ are equilibrium points for the flow: \begin{equation} N_\pm := (x_\pm,n_\pm),\qquad S_\pm := (x_\pm, s_\pm). \end{equation} We shall further assume that the flow does not have any non-wandering points other than the above four equilibria. The following assumptions fix the character of the four equilibrium points: \begin{equation} f'(x_-^{})\geq 0,\quad f'(x_+^{})\leq 0,\quad f''(x_\pm) \ne 0 \end{equation} (where by $f'(x_\pm)$ we mean the left derivate at $x_+^{}$ and the right derivative at $x_-^{}$), and \begin{equation}\label{cond:hyp} D_yg(x_-^{},n_-^{})>0,\quad D_yg(x_-^{},s_-^{})<0,\quad D_yg(x_+^{},n_+^{})<0,\quad D_yg(x_+^{},s_+^{})>0, \end{equation} where $D_yg$ is the $y$-derivative of $g(x,y)$. Thus $N_-$ is a (source) node, $N_+$ a (sink) node, and $S_\pm$ are saddle points. These will be hyperbolic if $f'(x_\pm) \ne 0$, and non-hyperbolic (degenerate) otherwise. Later on, in order to have a well-defined notion of index for certain distinguished orbits on $\mathcal{C}$, we will also assume that the locations of the equilibria on the boundary of the cylinder are not arbitrary, but are subject to the single condition \begin{equation}\label{assump} s_-^{} - n_-^{} = n_+^{} - s_+^{} \qquad(\mbox{mod}\ 2\pi) \end{equation} (This is a condition on $g(x,y)$. Although we will not pursue this approach here, under this condition \refeq{eq:flow} can be viewed as a flow with two equilibrium points on a {\em 2-torus}.) For a point $p\in\mathcal{C}$, let $\mathcal{O}(p)$ denote the flow orbit through $p$. Since $\mathcal{C}$ is compact, all orbits are complete, meaning they exist for all $s\in {\mathbb R}$, and since the flow is autonomous, two orbits are either disjoint or they coincide. The orbit of an equilibrium point consists of only one point, namely the equilibrium itself. The $\omega$-limit of any other orbit in $\mathcal{C}$ can be either $N_+$ or $S_+$, and the $\alpha$-limit likewise can only be either $N_-$ or $S_-$. All these facts are easy consequences of the existence and uniqueness theorem for ODEs. \subsubsection{Connecting orbits and corridors} Given a flow on $\mathcal{C}$ as in the above, there are two distinguished orbits in the interior of the cylinder: Let ${\mathcal{W}}^-$ denote the unique orbit of the flow whose $\alpha$-limit point is the saddle $S_-$, and let ${\mathcal{W}}^+$ denote the unique orbit whose $\omega$-limit point is the saddle $S_+$. In the hyperbolic case ($f'(x_\pm)\ne 0$) the uniqueness is immediate because ${\mathcal{W}}^-$ is the unstable manifold of $S_-$ and ${\mathcal{W}}^+$ is the stable manifold of $S_+$. In the non-hyperbolic case $f'(x_\pm)= 0$ the orbits ${\mathcal{W}}^\pm$ are {\em center} manifolds for the corresponding saddle-nodes $S_\pm$. Recall that center manifolds may be non-unique, but in our case the uniqueness is assured because the equilibrium points are on the boundary of the domain, so the relevant part of the center manifolds are on the ``saddle side'' of the equilibrium, and not on the ``node side'' (see Figure~\ref{fig:sn}). \begin{figure}[ht] \begin{center} \includegraphics[scale=0.3]{KTZ_zGKN_I_saddlenodeleft} \end{center} \caption{\label{fig:sn} Flow near a saddle-node. The node part lies outside of the domain of concern.} \end{figure} If ${\mathcal{W}}^+$ and ${\mathcal{W}}^-$ intersect, they must coincide, and the resulting orbit will connect the two saddle points, i.e. it will be the saddles connector we are after. Let us therefore assume that they are disjoint. The $\omega$-limit of ${\mathcal{W}}^-$ must then necessarily be $N_+$, and the $\alpha$-limit of ${\mathcal{W}}^+$ must be $N_-$. On the other hand the assumptions we have made about the flow imply that there are also two orbits of the flow on the left boundary of the cylinder connecting $N_-$ with $S_-$, call them $(N_-S_-)_\pm$, with $+$ denoting the counterclockwise one (when viewed from a point on the cylinder's axis and to the left of the cylinder), and similarly two joining $S_+$ with $N_+$, called $(S_+N_+)_\pm$. Consider therefore the following collection of six heteroclinic orbits \begin{equation} {\mathcal H} := \{ (N_-S_-)_\pm, {\mathcal{W}}^\pm, (S_+N_+)_\pm\}. \end{equation} The cylinder $\mathcal{C}$ is divided into two invariant regions ${\mathcal K}_1$ and ${\mathcal K}_2$, called {\em corridors}, by these orbits: $\mathcal{C} = {\mathcal K}_1 \cup {\mathcal H} \cup {\mathcal K}_2$. We would like to distinguish one of these two corridors. We do so as follows: Consider the lifting of the flow to the universal cover $\bar{\mathcal{C}}$. Let $\widetilde{S}_-$ denote the unique copy of the node $S_-$ that lies in the fundamental domain $\widetilde{\mathcal{C}}$, and let $\widetilde{\mathcal{W}}^-$ be the unique orbit in $\bar{\mathcal{C}}$ whose $\alpha$-limit point is $\widetilde{S}_-$. The $\omega$-limit point of this orbit is thus some copy of the node $N_+$, call it$\bar{N}_+$, which has coordinates $(x_+^{},n_+^{}-2\pi k_+)$ for some $k_+\in {\mathbb Z}$. Similarly, let $\widetilde{S}_+$ denote the unique point in the preimage of $S_+$ under the covering map that lies in the fundamental domain $\widetilde{\mathcal{C}}$ and let $\widetilde{\mathcal{W}}^+$ denote the unique orbit whose $\omega$-limit point is this $\widetilde{S}_+$. Let $\bar{N}_- = (x_-^{},n_-^{} +2\pi k_-)$, $k_-\in {\mathbb Z}$ be the $\alpha$-limit point of $\widetilde{\mathcal{W}}^+$. By definition the corridor ${\mathcal K}_1$ is the open domain in $\bar{\mathcal{C}}$ whose boundary contains the two orbits $\widetilde{\mathcal{W}}^-$ and $\widetilde{\mathcal{W}}^+$. We note that in $\bar{\mathcal{C}}$ only one of the two corridors will have {\em both} of these orbits on its boundary, so this is the distinguishing feature of ${\mathcal K}_1$. We orient the boundary of ${\mathcal K}_1$ (which is a closed simple curve) in such a way that the orientation induced on $\widetilde{\mathcal{W}}^-$ coincides with the direction of the flow on that orbit. \begin{figure}[ht] \begin{center} \font\thinlinefont=cmr5 \begingroup\makeatletter\ifx\SetFigFont\undefined% \gdef\SetFigFont#1#2#3#4#5{% \reset@font\fontsize{#1}{#2pt}% \fontfamily{#3}\fontseries{#4}\fontshape{#5}% \selectfont}% \fi\endgroup% \mbox{\beginpicture \setcoordinatesystem units <0.50000cm,0.50000cm> \unitlength=0.50000cm \linethickness=1pt \setplotsymbol ({\makebox(0,0)[l]{\tencirc\symbol{'160}}}) \setshadesymbol ({\thinlinefont .}) \setlinear \linethickness= 0.500pt \setplotsymbol ({\thinlinefont .}) {\color[rgb]{0,0,0}\ellipticalarc axes ratio 0.953:2.540 360 degrees from 6.032 16.510 center at 5.080 16.510 }% \linethickness= 0.500pt \setplotsymbol ({\thinlinefont .}) {\color[rgb]{0,0,0}\putrule from 5.080 13.970 to 12.700 13.970 }% \linethickness= 0.500pt \setplotsymbol ({\thinlinefont .}) {\color[rgb]{0,0,0}\putrule from 5.080 19.050 to 12.700 19.050 }% \linethickness= 0.500pt \setplotsymbol ({\thinlinefont .}) {\color[rgb]{0,0,0}\plot 5.397 14.287 5.715 13.970 / }% \linethickness= 0.500pt \setplotsymbol ({\thinlinefont .}) {\color[rgb]{0,0,0}\plot 5.715 14.605 6.350 13.970 / }% \linethickness= 0.500pt \setplotsymbol ({\thinlinefont .}) {\color[rgb]{0,0,0}\plot 7.010 13.966 5.804 15.045 / }% \linethickness= 0.500pt \setplotsymbol ({\thinlinefont .}) {\color[rgb]{0,0,0}\plot 5.994 15.553 7.391 14.199 / }% \linethickness= 0.500pt \setplotsymbol ({\thinlinefont .}) {\color[rgb]{0,0,0}\plot 6.058 16.104 7.942 14.389 / }% \linethickness= 0.500pt \setplotsymbol ({\thinlinefont .}) {\color[rgb]{0,0,0}\plot 6.037 16.739 8.175 14.685 / }% \linethickness= 0.500pt \setplotsymbol ({\thinlinefont .}) {\color[rgb]{0,0,0}\plot 7.645 15.828 8.492 15.024 / }% \linethickness= 0.500pt \setplotsymbol ({\thinlinefont .}) {\color[rgb]{0,0,0}\plot 8.407 15.680 8.894 15.257 / }% \linethickness= 0.500pt \setplotsymbol ({\thinlinefont .}) {\color[rgb]{0,0,0}\plot 8.894 15.828 9.318 15.405 / }% \linethickness= 0.500pt \setplotsymbol ({\thinlinefont .}) {\color[rgb]{0,0,0}\plot 9.275 16.082 9.974 15.447 / }% \linethickness= 0.500pt \setplotsymbol ({\thinlinefont .}) {\color[rgb]{0,0,0}\plot 9.529 16.463 12.133 13.923 / }% \linethickness= 0.500pt \setplotsymbol ({\thinlinefont .}) {\color[rgb]{0,0,0}\plot 10.524 14.855 11.540 13.902 / }% \linethickness= 0.500pt \setplotsymbol ({\thinlinefont .}) {\color[rgb]{0,0,0}\plot 9.847 16.739 12.725 13.987 / }% \linethickness= 0.500pt \setplotsymbol ({\thinlinefont .}) {\color[rgb]{0,0,0}\plot 10.291 17.035 12.979 14.326 / }% \linethickness= 0.500pt \setplotsymbol ({\thinlinefont .}) {\color[rgb]{0,0,0}\plot 10.736 17.204 13.233 14.707 / }% \linethickness= 0.500pt \setplotsymbol ({\thinlinefont .}) {\color[rgb]{0,0,0}\plot 11.223 17.289 13.382 15.151 / }% \linethickness= 0.500pt \setplotsymbol ({\thinlinefont .}) {\color[rgb]{0,0,0}\plot 12.048 17.120 13.318 15.828 / }% \linethickness= 0.500pt \setplotsymbol ({\thinlinefont .}) {\color[rgb]{0,0,0}\plot 4.386 17.691 4.216 17.479 / \putrule from 4.216 17.479 to 4.153 17.479 }% \linethickness= 0.500pt \setplotsymbol ({\thinlinefont .}) {\color[rgb]{0,0,0}\plot 4.534 17.141 4.153 16.760 / }% \linethickness= 0.500pt \setplotsymbol ({\thinlinefont .}) {\color[rgb]{0,0,0}\plot 4.682 16.675 4.132 16.125 / }% \linethickness= 0.500pt \setplotsymbol ({\thinlinefont .}) {\color[rgb]{0,0,0}\plot 4.873 16.188 4.153 15.617 / }% \linethickness= 0.500pt \setplotsymbol ({\thinlinefont .}) {\color[rgb]{0,0,0}\plot 5.021 15.807 4.301 15.088 / }% \linethickness= 0.500pt \setplotsymbol ({\thinlinefont .}) {\color[rgb]{0,0,0}\plot 5.254 15.405 4.492 14.601 / }% \linethickness= 0.500pt \setplotsymbol ({\thinlinefont .}) {\color[rgb]{0,0,0}\plot 5.508 14.939 4.640 14.156 / }% \linethickness= 0.500pt \setplotsymbol ({\thinlinefont .}) {\color[rgb]{0,0,0}\plot 5.719 14.601 5.063 13.966 / }% \linethickness= 0.500pt \setplotsymbol ({\thinlinefont .}) {\color[rgb]{0,0,0}\plot 6.037 14.326 5.698 13.966 / }% \linethickness= 0.500pt \setplotsymbol ({\thinlinefont .}) {\color[rgb]{0,0,0}\plot 6.460 14.114 6.354 13.966 / }% \linethickness= 0.500pt \setplotsymbol ({\thinlinefont .}) {\color[rgb]{0,0,0}\plot 12.112 14.601 11.434 13.966 / }% \linethickness= 0.500pt \setplotsymbol ({\thinlinefont .}) {\color[rgb]{0,0,0}\plot 12.344 14.220 12.112 13.987 / }% \linethickness= 0.500pt \setplotsymbol ({\thinlinefont .}) {\color[rgb]{0,0,0}\plot 11.921 15.066 11.223 14.410 / }% \linethickness= 0.500pt \setplotsymbol ({\thinlinefont .}) {\color[rgb]{0,0,0}\plot 11.836 15.596 11.434 15.257 / }% \linethickness= 0.500pt \setplotsymbol ({\thinlinefont .}) {\color[rgb]{0,0,0}\plot 11.773 16.209 11.498 15.892 / }% \linethickness= 0.500pt \setplotsymbol ({\thinlinefont .}) {\color[rgb]{0,0,0}\plot 11.773 16.781 11.604 16.590 / }% \linethickness= 0.500pt \setplotsymbol ({\thinlinefont .}) {\color[rgb]{0,0,0}\plot 11.985 17.797 11.667 17.416 / }% \linethickness= 0.500pt \setplotsymbol ({\thinlinefont .}) {\color[rgb]{0,0,0}\plot 12.090 18.411 11.709 17.987 / }% \linethickness= 0.500pt \setplotsymbol ({\thinlinefont .}) {\color[rgb]{0,0,0}\plot 12.090 19.025 11.836 18.792 / }% \linethickness= 0.500pt \setplotsymbol ({\thinlinefont .}) {\color[rgb]{0,0,0}\plot 12.493 19.046 13.022 18.601 / }% \linethickness= 0.500pt \setplotsymbol ({\thinlinefont .}) {\color[rgb]{0,0,0}\plot 12.958 18.347 13.360 17.966 / }% \linethickness=1pt \setplotsymbol ({\makebox(0,0)[l]{\tencirc\symbol{'160}}}) {\color[rgb]{0,0,0}\plot 5.715 18.415 5.717 18.411 / \plot 5.717 18.411 5.721 18.400 / \plot 5.721 18.400 5.730 18.379 / \plot 5.730 18.379 5.743 18.349 / \plot 5.743 18.349 5.762 18.309 / \plot 5.762 18.309 5.783 18.260 / \plot 5.783 18.260 5.808 18.201 / \plot 5.808 18.201 5.836 18.138 / \plot 5.836 18.138 5.865 18.070 / \plot 5.865 18.070 5.897 18.000 / \plot 5.897 18.000 5.927 17.930 / \plot 5.927 17.930 5.956 17.863 / \plot 5.956 17.863 5.986 17.799 / \plot 5.986 17.799 6.013 17.738 / \plot 6.013 17.738 6.039 17.678 / \plot 6.039 17.678 6.064 17.623 / \plot 6.064 17.623 6.088 17.573 / \plot 6.088 17.573 6.111 17.524 / \plot 6.111 17.524 6.134 17.477 / \plot 6.134 17.477 6.155 17.433 / \plot 6.155 17.433 6.176 17.388 / \plot 6.176 17.388 6.198 17.346 / \plot 6.198 17.346 6.219 17.304 / \plot 6.219 17.304 6.240 17.261 / \plot 6.240 17.261 6.261 17.219 / \plot 6.261 17.219 6.284 17.175 / \plot 6.284 17.175 6.308 17.130 / \plot 6.308 17.130 6.331 17.086 / \plot 6.331 17.086 6.356 17.041 / \plot 6.356 17.041 6.382 16.995 / \plot 6.382 16.995 6.409 16.948 / \plot 6.409 16.948 6.439 16.902 / \plot 6.439 16.902 6.469 16.853 / \plot 6.469 16.853 6.498 16.806 / \plot 6.498 16.806 6.530 16.760 / \plot 6.530 16.760 6.564 16.713 / \plot 6.564 16.713 6.598 16.667 / \plot 6.598 16.667 6.632 16.620 / \plot 6.632 16.620 6.665 16.578 / \plot 6.665 16.578 6.699 16.533 / \plot 6.699 16.533 6.735 16.493 / \plot 6.735 16.493 6.771 16.453 / \plot 6.771 16.453 6.807 16.413 / \plot 6.807 16.413 6.845 16.377 / \plot 6.845 16.377 6.881 16.341 / \plot 6.881 16.341 6.919 16.305 / \plot 6.919 16.305 6.960 16.271 / \plot 6.960 16.271 6.998 16.239 / \plot 6.998 16.239 7.040 16.207 / \plot 7.040 16.207 7.082 16.173 / \plot 7.082 16.173 7.129 16.142 / \plot 7.129 16.142 7.176 16.112 / \plot 7.176 16.112 7.224 16.080 / \plot 7.224 16.080 7.275 16.051 / \plot 7.275 16.051 7.326 16.019 / \plot 7.326 16.019 7.379 15.989 / \plot 7.379 15.989 7.434 15.962 / \plot 7.434 15.962 7.489 15.934 / \plot 7.489 15.934 7.544 15.909 / \plot 7.544 15.909 7.601 15.883 / \plot 7.601 15.883 7.656 15.860 / \plot 7.656 15.860 7.711 15.837 / \plot 7.711 15.837 7.766 15.818 / \plot 7.766 15.818 7.819 15.799 / \plot 7.819 15.799 7.870 15.782 / \plot 7.870 15.782 7.921 15.767 / \plot 7.921 15.767 7.969 15.752 / \plot 7.969 15.752 8.016 15.742 / \plot 8.016 15.742 8.062 15.731 / \plot 8.062 15.731 8.107 15.723 / \plot 8.107 15.723 8.149 15.716 / \plot 8.149 15.716 8.200 15.710 / \plot 8.200 15.710 8.249 15.706 / \plot 8.249 15.706 8.297 15.704 / \putrule from 8.297 15.704 to 8.348 15.704 \plot 8.348 15.704 8.397 15.708 / \plot 8.397 15.708 8.446 15.712 / \plot 8.446 15.712 8.494 15.718 / \plot 8.494 15.718 8.543 15.729 / \plot 8.543 15.729 8.592 15.742 / \plot 8.592 15.742 8.640 15.756 / \plot 8.640 15.756 8.689 15.773 / \plot 8.689 15.773 8.735 15.792 / \plot 8.735 15.792 8.782 15.814 / \plot 8.782 15.814 8.827 15.837 / \plot 8.827 15.837 8.871 15.862 / \plot 8.871 15.862 8.913 15.888 / \plot 8.913 15.888 8.954 15.915 / \plot 8.954 15.915 8.996 15.945 / \plot 8.996 15.945 9.036 15.974 / \plot 9.036 15.974 9.076 16.006 / \plot 9.076 16.006 9.108 16.036 / \plot 9.108 16.036 9.142 16.066 / \plot 9.142 16.066 9.178 16.095 / \plot 9.178 16.095 9.214 16.127 / \plot 9.214 16.127 9.250 16.161 / \plot 9.250 16.161 9.288 16.197 / \plot 9.288 16.197 9.326 16.235 / \plot 9.326 16.235 9.366 16.273 / \plot 9.366 16.273 9.409 16.311 / \plot 9.409 16.311 9.451 16.351 / \plot 9.451 16.351 9.493 16.394 / \plot 9.493 16.394 9.536 16.434 / \plot 9.536 16.434 9.580 16.476 / \plot 9.580 16.476 9.624 16.516 / \plot 9.624 16.516 9.669 16.559 / \plot 9.669 16.559 9.711 16.599 / \plot 9.711 16.599 9.756 16.639 / \plot 9.756 16.639 9.800 16.677 / \plot 9.800 16.677 9.842 16.713 / \plot 9.842 16.713 9.885 16.749 / \plot 9.885 16.749 9.927 16.785 / \plot 9.927 16.785 9.970 16.817 / \plot 9.970 16.817 10.012 16.849 / \plot 10.012 16.849 10.054 16.880 / \plot 10.054 16.880 10.097 16.910 / \plot 10.097 16.910 10.141 16.940 / \plot 10.141 16.940 10.185 16.967 / \plot 10.185 16.967 10.230 16.995 / \plot 10.230 16.995 10.279 17.022 / \plot 10.279 17.022 10.327 17.050 / \plot 10.327 17.050 10.376 17.075 / \plot 10.376 17.075 10.427 17.101 / \plot 10.427 17.101 10.480 17.124 / \plot 10.480 17.124 10.533 17.147 / \plot 10.533 17.147 10.585 17.170 / \plot 10.585 17.170 10.640 17.192 / \plot 10.640 17.192 10.693 17.211 / \plot 10.693 17.211 10.748 17.230 / \plot 10.748 17.230 10.801 17.247 / \plot 10.801 17.247 10.854 17.261 / \plot 10.854 17.261 10.907 17.276 / \plot 10.907 17.276 10.958 17.287 / \plot 10.958 17.287 11.009 17.300 / \plot 11.009 17.300 11.057 17.308 / \plot 11.057 17.308 11.106 17.314 / \plot 11.106 17.314 11.153 17.321 / \plot 11.153 17.321 11.199 17.327 / \plot 11.199 17.327 11.246 17.329 / \plot 11.246 17.329 11.295 17.333 / \putrule from 11.295 17.333 to 11.343 17.333 \putrule from 11.343 17.333 to 11.394 17.333 \plot 11.394 17.333 11.445 17.331 / \plot 11.445 17.331 11.496 17.329 / \plot 11.496 17.329 11.549 17.323 / \plot 11.549 17.323 11.601 17.316 / \plot 11.601 17.316 11.654 17.308 / \plot 11.654 17.308 11.709 17.300 / \plot 11.709 17.300 11.762 17.287 / \plot 11.762 17.287 11.815 17.274 / \plot 11.815 17.274 11.868 17.259 / \plot 11.868 17.259 11.921 17.244 / \plot 11.921 17.244 11.972 17.228 / \plot 11.972 17.228 12.023 17.211 / \plot 12.023 17.211 12.071 17.192 / \plot 12.071 17.192 12.118 17.173 / \plot 12.118 17.173 12.162 17.153 / \plot 12.162 17.153 12.207 17.132 / \plot 12.207 17.132 12.249 17.111 / \plot 12.249 17.111 12.289 17.088 / \plot 12.289 17.088 12.330 17.065 / \plot 12.330 17.065 12.370 17.041 / \plot 12.370 17.041 12.408 17.018 / \plot 12.408 17.018 12.448 16.990 / \plot 12.448 16.990 12.486 16.965 / \plot 12.486 16.965 12.526 16.935 / \plot 12.526 16.935 12.565 16.906 / \plot 12.565 16.906 12.605 16.874 / \plot 12.605 16.874 12.645 16.842 / \plot 12.645 16.842 12.683 16.808 / \plot 12.683 16.808 12.723 16.772 / \plot 12.723 16.772 12.761 16.736 / \plot 12.761 16.736 12.797 16.701 / \plot 12.797 16.701 12.835 16.665 / \plot 12.835 16.665 12.869 16.629 / \plot 12.869 16.629 12.903 16.593 / \plot 12.903 16.593 12.935 16.557 / \plot 12.935 16.557 12.967 16.521 / \plot 12.967 16.521 12.994 16.487 / \plot 12.994 16.487 13.022 16.451 / \plot 13.022 16.451 13.049 16.417 / \plot 13.049 16.417 13.073 16.385 / \plot 13.073 16.385 13.098 16.351 / \plot 13.098 16.351 13.121 16.315 / \plot 13.121 16.315 13.147 16.277 / \plot 13.147 16.277 13.170 16.241 / \plot 13.170 16.241 13.193 16.201 / \plot 13.193 16.201 13.216 16.159 / \plot 13.216 16.159 13.240 16.114 / \plot 13.240 16.114 13.263 16.068 / \plot 13.263 16.068 13.288 16.017 / \plot 13.288 16.017 13.316 15.962 / \plot 13.316 15.962 13.341 15.903 / \plot 13.341 15.903 13.369 15.843 / \plot 13.369 15.843 13.394 15.784 / \plot 13.394 15.784 13.420 15.729 / \plot 13.420 15.729 13.443 15.676 / \plot 13.443 15.676 13.462 15.634 / \plot 13.462 15.634 13.477 15.600 / \plot 13.477 15.600 13.485 15.577 / \plot 13.485 15.577 13.492 15.564 / \plot 13.492 15.564 13.494 15.558 / }% \linethickness=1pt \setplotsymbol ({\makebox(0,0)[l]{\tencirc\symbol{'160}}}) {\color[rgb]{0,0,0}\plot 6.985 13.970 6.991 13.972 / \plot 6.991 13.972 7.004 13.978 / \plot 7.004 13.978 7.027 13.987 / \plot 7.027 13.987 7.059 14.002 / \plot 7.059 14.002 7.104 14.021 / \plot 7.104 14.021 7.154 14.044 / \plot 7.154 14.044 7.209 14.069 / \plot 7.209 14.069 7.269 14.095 / \plot 7.269 14.095 7.326 14.122 / \plot 7.326 14.122 7.383 14.148 / \plot 7.383 14.148 7.436 14.175 / \plot 7.436 14.175 7.487 14.201 / \plot 7.487 14.201 7.531 14.224 / \plot 7.531 14.224 7.573 14.247 / \plot 7.573 14.247 7.614 14.271 / \plot 7.614 14.271 7.650 14.294 / \plot 7.650 14.294 7.686 14.317 / \plot 7.686 14.317 7.719 14.343 / \plot 7.719 14.343 7.753 14.366 / \plot 7.753 14.366 7.785 14.393 / \plot 7.785 14.393 7.817 14.421 / \plot 7.817 14.421 7.851 14.448 / \plot 7.851 14.448 7.885 14.478 / \plot 7.885 14.478 7.916 14.510 / \plot 7.916 14.510 7.950 14.544 / \plot 7.950 14.544 7.986 14.575 / \plot 7.986 14.575 8.020 14.611 / \plot 8.020 14.611 8.054 14.645 / \plot 8.054 14.645 8.090 14.681 / \plot 8.090 14.681 8.124 14.715 / \plot 8.124 14.715 8.156 14.749 / \plot 8.156 14.749 8.189 14.783 / \plot 8.189 14.783 8.221 14.815 / \plot 8.221 14.815 8.251 14.846 / \plot 8.251 14.846 8.280 14.876 / \plot 8.280 14.876 8.308 14.903 / \plot 8.308 14.903 8.335 14.929 / \plot 8.335 14.929 8.361 14.952 / \plot 8.361 14.952 8.388 14.975 / \plot 8.388 14.975 8.418 15.001 / \plot 8.418 15.001 8.450 15.026 / \plot 8.450 15.026 8.484 15.050 / \plot 8.484 15.050 8.515 15.073 / \plot 8.515 15.073 8.551 15.094 / \plot 8.551 15.094 8.585 15.115 / \plot 8.585 15.115 8.621 15.134 / \plot 8.621 15.134 8.659 15.153 / \plot 8.659 15.153 8.697 15.172 / \plot 8.697 15.172 8.735 15.189 / \plot 8.735 15.189 8.774 15.208 / \plot 8.774 15.208 8.812 15.225 / \plot 8.812 15.225 8.852 15.242 / \plot 8.852 15.242 8.890 15.259 / \plot 8.890 15.259 8.928 15.276 / \plot 8.928 15.276 8.970 15.293 / \plot 8.970 15.293 9.002 15.308 / \plot 9.002 15.308 9.038 15.323 / \plot 9.038 15.323 9.074 15.339 / \plot 9.074 15.339 9.112 15.356 / \plot 9.112 15.356 9.150 15.373 / \plot 9.150 15.373 9.193 15.392 / \plot 9.193 15.392 9.235 15.409 / \plot 9.235 15.409 9.279 15.426 / \plot 9.279 15.426 9.324 15.443 / \plot 9.324 15.443 9.370 15.460 / \plot 9.370 15.460 9.415 15.475 / \plot 9.415 15.475 9.462 15.490 / \plot 9.462 15.490 9.506 15.500 / \plot 9.506 15.500 9.550 15.511 / \plot 9.550 15.511 9.593 15.519 / \plot 9.593 15.519 9.635 15.528 / \plot 9.635 15.528 9.675 15.532 / \plot 9.675 15.532 9.716 15.534 / \putrule from 9.716 15.534 to 9.754 15.534 \plot 9.754 15.534 9.790 15.530 / \plot 9.790 15.530 9.826 15.526 / \plot 9.826 15.526 9.862 15.519 / \plot 9.862 15.519 9.898 15.509 / \plot 9.898 15.509 9.934 15.496 / \plot 9.934 15.496 9.967 15.481 / \plot 9.967 15.481 10.003 15.464 / \plot 10.003 15.464 10.039 15.443 / \plot 10.039 15.443 10.073 15.420 / \plot 10.073 15.420 10.107 15.397 / \plot 10.107 15.397 10.141 15.369 / \plot 10.141 15.369 10.175 15.342 / \plot 10.175 15.342 10.204 15.312 / \plot 10.204 15.312 10.236 15.280 / \plot 10.236 15.280 10.264 15.248 / \plot 10.264 15.248 10.289 15.217 / \plot 10.289 15.217 10.315 15.185 / \plot 10.315 15.185 10.338 15.153 / \plot 10.338 15.153 10.359 15.119 / \plot 10.359 15.119 10.380 15.088 / \plot 10.380 15.088 10.399 15.054 / \plot 10.399 15.054 10.418 15.018 / \plot 10.418 15.018 10.437 14.980 / \plot 10.437 14.980 10.454 14.942 / \plot 10.454 14.942 10.471 14.901 / \plot 10.471 14.901 10.488 14.861 / \plot 10.488 14.861 10.505 14.819 / \plot 10.505 14.819 10.520 14.776 / \plot 10.520 14.776 10.535 14.734 / \plot 10.535 14.734 10.549 14.692 / \plot 10.549 14.692 10.564 14.649 / \plot 10.564 14.649 10.577 14.609 / \plot 10.577 14.609 10.590 14.569 / \plot 10.590 14.569 10.602 14.531 / \plot 10.602 14.531 10.615 14.493 / \plot 10.615 14.493 10.628 14.459 / \plot 10.628 14.459 10.638 14.427 / \plot 10.638 14.427 10.651 14.395 / \plot 10.651 14.395 10.664 14.366 / \plot 10.664 14.366 10.679 14.332 / \plot 10.679 14.332 10.696 14.298 / \plot 10.696 14.298 10.715 14.264 / \plot 10.715 14.264 10.736 14.230 / \plot 10.736 14.230 10.761 14.196 / \plot 10.761 14.196 10.789 14.161 / \plot 10.789 14.161 10.820 14.122 / \plot 10.820 14.122 10.852 14.084 / \plot 10.852 14.084 10.884 14.046 / \plot 10.884 14.046 10.914 14.014 / \plot 10.914 14.014 10.935 13.991 / \plot 10.935 13.991 10.947 13.976 / \plot 10.947 13.976 10.954 13.970 / }% \linethickness=1pt \setplotsymbol ({\makebox(0,0)[l]{\tencirc\symbol{'160}}}) {\color[rgb]{0,0,0}\plot 4.286 18.098 4.288 18.093 / \plot 4.288 18.093 4.290 18.083 / \plot 4.290 18.083 4.295 18.066 / \plot 4.295 18.066 4.301 18.036 / \plot 4.301 18.036 4.312 17.998 / \plot 4.312 17.998 4.326 17.947 / \plot 4.326 17.947 4.341 17.886 / \plot 4.341 17.886 4.360 17.814 / \plot 4.360 17.814 4.381 17.733 / \plot 4.381 17.733 4.405 17.649 / \plot 4.405 17.649 4.430 17.558 / \plot 4.430 17.558 4.456 17.465 / \plot 4.456 17.465 4.481 17.369 / \plot 4.481 17.369 4.508 17.276 / \plot 4.508 17.276 4.534 17.185 / \plot 4.534 17.185 4.559 17.096 / \plot 4.559 17.096 4.583 17.012 / \plot 4.583 17.012 4.606 16.931 / \plot 4.606 16.931 4.629 16.855 / \plot 4.629 16.855 4.652 16.781 / \plot 4.652 16.781 4.674 16.711 / \plot 4.674 16.711 4.695 16.645 / \plot 4.695 16.645 4.716 16.582 / \plot 4.716 16.582 4.737 16.523 / \plot 4.737 16.523 4.758 16.466 / \plot 4.758 16.466 4.777 16.408 / \plot 4.777 16.408 4.798 16.353 / \plot 4.798 16.353 4.820 16.298 / \plot 4.820 16.298 4.843 16.245 / \plot 4.843 16.245 4.866 16.188 / \plot 4.866 16.188 4.889 16.131 / \plot 4.889 16.131 4.915 16.074 / \plot 4.915 16.074 4.942 16.017 / \plot 4.942 16.017 4.970 15.958 / \plot 4.970 15.958 5.000 15.896 / \plot 5.000 15.896 5.031 15.833 / \plot 5.031 15.833 5.063 15.767 / \plot 5.063 15.767 5.099 15.699 / \plot 5.099 15.699 5.137 15.627 / \plot 5.137 15.627 5.177 15.551 / \plot 5.177 15.551 5.222 15.473 / \plot 5.222 15.473 5.266 15.390 / \plot 5.266 15.390 5.313 15.306 / \plot 5.313 15.306 5.362 15.219 / \plot 5.362 15.219 5.410 15.132 / \plot 5.410 15.132 5.459 15.047 / \plot 5.459 15.047 5.508 14.965 / \plot 5.508 14.965 5.550 14.889 / \plot 5.550 14.889 5.592 14.819 / \plot 5.592 14.819 5.626 14.757 / \plot 5.626 14.757 5.656 14.707 / \plot 5.656 14.707 5.679 14.666 / \plot 5.679 14.666 5.696 14.639 / \plot 5.696 14.639 5.707 14.620 / \plot 5.707 14.620 5.713 14.609 / \plot 5.713 14.609 5.715 14.605 / }% \linethickness= 0.500pt \setplotsymbol ({\thinlinefont .}) {\color[rgb]{0,0,0}\plot 5.715 14.605 5.721 14.601 / \plot 5.721 14.601 5.736 14.590 / \plot 5.736 14.590 5.759 14.573 / \plot 5.759 14.573 5.793 14.548 / \plot 5.793 14.548 5.836 14.516 / \plot 5.836 14.516 5.884 14.480 / \plot 5.884 14.480 5.935 14.444 / \plot 5.935 14.444 5.988 14.406 / \plot 5.988 14.406 6.037 14.370 / \plot 6.037 14.370 6.083 14.338 / \plot 6.083 14.338 6.126 14.309 / \plot 6.126 14.309 6.166 14.281 / \plot 6.166 14.281 6.202 14.258 / \plot 6.202 14.258 6.234 14.237 / \plot 6.234 14.237 6.265 14.216 / \plot 6.265 14.216 6.295 14.199 / \plot 6.295 14.199 6.325 14.182 / \plot 6.325 14.182 6.361 14.161 / \plot 6.361 14.161 6.397 14.141 / \plot 6.397 14.141 6.433 14.125 / \plot 6.433 14.125 6.469 14.108 / \plot 6.469 14.108 6.505 14.091 / \plot 6.505 14.091 6.541 14.074 / \plot 6.541 14.074 6.576 14.061 / \plot 6.576 14.061 6.612 14.048 / \plot 6.612 14.048 6.644 14.036 / \plot 6.644 14.036 6.674 14.025 / \plot 6.674 14.025 6.703 14.017 / \plot 6.703 14.017 6.729 14.008 / \plot 6.729 14.008 6.752 14.002 / \plot 6.752 14.002 6.773 13.995 / \plot 6.773 13.995 6.801 13.989 / \plot 6.801 13.989 6.826 13.985 / \plot 6.826 13.985 6.852 13.981 / \plot 6.852 13.981 6.879 13.976 / \plot 6.879 13.976 6.907 13.974 / \plot 6.907 13.974 6.934 13.972 / \putrule from 6.934 13.972 to 6.960 13.972 \plot 6.960 13.972 6.977 13.970 / \putrule from 6.977 13.970 to 6.983 13.970 \putrule from 6.983 13.970 to 6.985 13.970 }% \linethickness= 0.500pt \setplotsymbol ({\thinlinefont .}) {\color[rgb]{0,0,0}\plot 11.113 13.970 11.115 13.976 / \plot 11.115 13.976 11.119 13.989 / \plot 11.119 13.989 11.125 14.010 / \plot 11.125 14.010 11.136 14.044 / \plot 11.136 14.044 11.148 14.089 / \plot 11.148 14.089 11.165 14.144 / \plot 11.165 14.144 11.184 14.207 / \plot 11.184 14.207 11.206 14.275 / \plot 11.206 14.275 11.225 14.345 / \plot 11.225 14.345 11.246 14.417 / \plot 11.246 14.417 11.267 14.489 / \plot 11.267 14.489 11.286 14.556 / \plot 11.286 14.556 11.303 14.624 / \plot 11.303 14.624 11.318 14.688 / \plot 11.318 14.688 11.335 14.749 / \plot 11.335 14.749 11.347 14.810 / \plot 11.347 14.810 11.360 14.867 / \plot 11.360 14.867 11.373 14.927 / \plot 11.373 14.927 11.383 14.986 / \plot 11.383 14.986 11.394 15.045 / \plot 11.394 15.045 11.405 15.107 / \plot 11.405 15.107 11.411 15.157 / \plot 11.411 15.157 11.419 15.208 / \plot 11.419 15.208 11.426 15.263 / \plot 11.426 15.263 11.434 15.318 / \plot 11.434 15.318 11.441 15.375 / \plot 11.441 15.375 11.449 15.433 / \plot 11.449 15.433 11.455 15.494 / \plot 11.455 15.494 11.464 15.558 / \plot 11.464 15.558 11.470 15.623 / \plot 11.470 15.623 11.477 15.689 / \plot 11.477 15.689 11.485 15.756 / \plot 11.485 15.756 11.491 15.824 / \plot 11.491 15.824 11.498 15.894 / \plot 11.498 15.894 11.506 15.966 / \plot 11.506 15.966 11.513 16.036 / \plot 11.513 16.036 11.519 16.108 / \plot 11.519 16.108 11.525 16.178 / \plot 11.525 16.178 11.532 16.248 / \plot 11.532 16.248 11.538 16.317 / \plot 11.538 16.317 11.544 16.385 / \plot 11.544 16.385 11.551 16.451 / \plot 11.551 16.451 11.557 16.516 / \plot 11.557 16.516 11.561 16.580 / \plot 11.561 16.580 11.568 16.643 / \plot 11.568 16.643 11.574 16.705 / \plot 11.574 16.705 11.578 16.764 / \plot 11.578 16.764 11.585 16.823 / \plot 11.585 16.823 11.589 16.880 / \plot 11.589 16.880 11.595 16.946 / \plot 11.595 16.946 11.601 17.012 / \plot 11.601 17.012 11.608 17.077 / \plot 11.608 17.077 11.614 17.143 / \plot 11.614 17.143 11.620 17.209 / \plot 11.620 17.209 11.627 17.274 / \plot 11.627 17.274 11.633 17.342 / \plot 11.633 17.342 11.640 17.407 / \plot 11.640 17.407 11.646 17.473 / \plot 11.646 17.473 11.654 17.537 / \plot 11.654 17.537 11.661 17.600 / \plot 11.661 17.600 11.669 17.664 / \plot 11.669 17.664 11.676 17.725 / \plot 11.676 17.725 11.682 17.784 / \plot 11.682 17.784 11.690 17.844 / \plot 11.690 17.844 11.697 17.899 / \plot 11.697 17.899 11.703 17.951 / \plot 11.703 17.951 11.709 18.002 / \plot 11.709 18.002 11.716 18.051 / \plot 11.716 18.051 11.722 18.095 / \plot 11.722 18.095 11.728 18.140 / \plot 11.728 18.140 11.735 18.180 / \plot 11.735 18.180 11.741 18.218 / \plot 11.741 18.218 11.748 18.256 / \plot 11.748 18.256 11.756 18.309 / \plot 11.756 18.309 11.767 18.360 / \plot 11.767 18.360 11.775 18.409 / \plot 11.775 18.409 11.786 18.455 / \plot 11.786 18.455 11.794 18.500 / \plot 11.794 18.500 11.805 18.544 / \plot 11.805 18.544 11.813 18.584 / \plot 11.813 18.584 11.824 18.625 / \plot 11.824 18.625 11.832 18.661 / \plot 11.832 18.661 11.841 18.694 / \plot 11.841 18.694 11.849 18.724 / \plot 11.849 18.724 11.858 18.752 / \plot 11.858 18.752 11.864 18.777 / \plot 11.864 18.777 11.870 18.798 / \plot 11.870 18.798 11.874 18.819 / \plot 11.874 18.819 11.881 18.838 / \plot 11.881 18.838 11.887 18.866 / \plot 11.887 18.866 11.891 18.891 / \plot 11.891 18.891 11.896 18.917 / \plot 11.896 18.917 11.900 18.944 / \plot 11.900 18.944 11.902 18.972 / \plot 11.902 18.972 11.904 18.999 / \putrule from 11.904 18.999 to 11.904 19.025 \plot 11.904 19.025 11.906 19.042 / \putrule from 11.906 19.042 to 11.906 19.048 \putrule from 11.906 19.048 to 11.906 19.050 }% \linethickness= 0.500pt \setplotsymbol ({\thinlinefont .}) {\color[rgb]{0,0,0}\putrule from 12.700 19.050 to 12.698 19.050 \putrule from 12.698 19.050 to 12.689 19.050 \plot 12.689 19.050 12.670 19.048 / \plot 12.670 19.048 12.639 19.044 / \plot 12.639 19.044 12.601 19.039 / \plot 12.601 19.039 12.562 19.031 / \plot 12.562 19.031 12.524 19.022 / \plot 12.524 19.022 12.490 19.012 / \plot 12.490 19.012 12.459 18.999 / \plot 12.459 18.999 12.431 18.984 / \plot 12.431 18.984 12.408 18.967 / \plot 12.408 18.967 12.383 18.944 / \plot 12.383 18.944 12.366 18.927 / \plot 12.366 18.927 12.351 18.908 / \plot 12.351 18.908 12.334 18.887 / \plot 12.334 18.887 12.317 18.862 / \plot 12.317 18.862 12.298 18.836 / \plot 12.298 18.836 12.281 18.804 / \plot 12.281 18.804 12.262 18.773 / \plot 12.262 18.773 12.243 18.737 / \plot 12.243 18.737 12.224 18.699 / \plot 12.224 18.699 12.205 18.658 / \plot 12.205 18.658 12.186 18.616 / \plot 12.186 18.616 12.167 18.574 / \plot 12.167 18.574 12.150 18.527 / \plot 12.150 18.527 12.131 18.481 / \plot 12.131 18.481 12.114 18.434 / \plot 12.114 18.434 12.097 18.383 / \plot 12.097 18.383 12.082 18.335 / \plot 12.082 18.335 12.065 18.282 / \plot 12.065 18.282 12.052 18.244 / \plot 12.052 18.244 12.040 18.201 / \plot 12.040 18.201 12.029 18.157 / \plot 12.029 18.157 12.016 18.112 / \plot 12.016 18.112 12.004 18.066 / \plot 12.004 18.066 11.991 18.017 / \plot 11.991 18.017 11.978 17.966 / \plot 11.978 17.966 11.966 17.913 / \plot 11.966 17.913 11.951 17.858 / \plot 11.951 17.858 11.938 17.803 / \plot 11.938 17.803 11.925 17.746 / \plot 11.925 17.746 11.913 17.689 / \plot 11.913 17.689 11.902 17.632 / \plot 11.902 17.632 11.889 17.575 / \plot 11.889 17.575 11.879 17.515 / \plot 11.879 17.515 11.866 17.458 / \plot 11.866 17.458 11.858 17.403 / \plot 11.858 17.403 11.847 17.348 / \plot 11.847 17.348 11.839 17.293 / \plot 11.839 17.293 11.830 17.240 / \plot 11.830 17.240 11.822 17.187 / \plot 11.822 17.187 11.813 17.137 / \plot 11.813 17.137 11.807 17.088 / \plot 11.807 17.088 11.800 17.039 / \plot 11.800 17.039 11.794 16.986 / \plot 11.794 16.986 11.788 16.933 / \plot 11.788 16.933 11.783 16.880 / \plot 11.783 16.880 11.779 16.828 / \plot 11.779 16.828 11.775 16.772 / \plot 11.775 16.772 11.771 16.717 / \plot 11.771 16.717 11.769 16.662 / \plot 11.769 16.662 11.764 16.605 / \putrule from 11.764 16.605 to 11.764 16.548 \plot 11.764 16.548 11.762 16.491 / \putrule from 11.762 16.491 to 11.762 16.434 \putrule from 11.762 16.434 to 11.762 16.377 \plot 11.762 16.377 11.764 16.320 / \putrule from 11.764 16.320 to 11.764 16.264 \plot 11.764 16.264 11.769 16.209 / \plot 11.769 16.209 11.771 16.154 / \plot 11.771 16.154 11.775 16.101 / \plot 11.775 16.101 11.779 16.049 / \plot 11.779 16.049 11.783 15.998 / \plot 11.783 15.998 11.788 15.947 / \plot 11.788 15.947 11.794 15.898 / \plot 11.794 15.898 11.800 15.847 / \plot 11.800 15.847 11.807 15.799 / \plot 11.807 15.799 11.815 15.748 / \plot 11.815 15.748 11.824 15.697 / \plot 11.824 15.697 11.832 15.644 / \plot 11.832 15.644 11.843 15.591 / \plot 11.843 15.591 11.851 15.536 / \plot 11.851 15.536 11.864 15.481 / \plot 11.864 15.481 11.874 15.424 / \plot 11.874 15.424 11.887 15.367 / \plot 11.887 15.367 11.900 15.312 / \plot 11.900 15.312 11.913 15.255 / \plot 11.913 15.255 11.927 15.200 / \plot 11.927 15.200 11.942 15.145 / \plot 11.942 15.145 11.955 15.092 / \plot 11.955 15.092 11.970 15.039 / \plot 11.970 15.039 11.982 14.990 / \plot 11.982 14.990 11.997 14.942 / \plot 11.997 14.942 12.012 14.897 / \plot 12.012 14.897 12.025 14.855 / \plot 12.025 14.855 12.037 14.812 / \plot 12.037 14.812 12.052 14.774 / \plot 12.052 14.774 12.065 14.736 / \plot 12.065 14.736 12.084 14.688 / \plot 12.084 14.688 12.101 14.641 / \plot 12.101 14.641 12.120 14.597 / \plot 12.120 14.597 12.141 14.552 / \plot 12.141 14.552 12.160 14.510 / \plot 12.160 14.510 12.181 14.470 / \plot 12.181 14.470 12.203 14.429 / \plot 12.203 14.429 12.224 14.393 / \plot 12.224 14.393 12.245 14.357 / \plot 12.245 14.357 12.266 14.326 / \plot 12.266 14.326 12.287 14.296 / \plot 12.287 14.296 12.306 14.268 / \plot 12.306 14.268 12.327 14.243 / \plot 12.327 14.243 12.347 14.222 / \plot 12.347 14.222 12.363 14.201 / \plot 12.363 14.201 12.383 14.182 / \plot 12.383 14.182 12.408 14.158 / \plot 12.408 14.158 12.431 14.137 / \plot 12.431 14.137 12.459 14.116 / \plot 12.459 14.116 12.490 14.095 / \plot 12.490 14.095 12.524 14.072 / \plot 12.524 14.072 12.562 14.048 / \plot 12.562 14.048 12.601 14.025 / \plot 12.601 14.025 12.639 14.004 / \plot 12.639 14.004 12.670 13.987 / \plot 12.670 13.987 12.689 13.976 / \plot 12.689 13.976 12.698 13.970 / \putrule from 12.698 13.970 to 12.700 13.970 }% \linethickness= 0.500pt \setplotsymbol ({\thinlinefont .}) {\color[rgb]{0,0,0}\plot 12.700 19.050 12.704 19.044 / \plot 12.704 19.044 12.711 19.033 / \plot 12.711 19.033 12.725 19.012 / \plot 12.725 19.012 12.747 18.980 / \plot 12.747 18.980 12.774 18.938 / \plot 12.774 18.938 12.806 18.887 / \plot 12.806 18.887 12.844 18.830 / \plot 12.844 18.830 12.884 18.766 / \plot 12.884 18.766 12.926 18.703 / \plot 12.926 18.703 12.967 18.639 / \plot 12.967 18.639 13.007 18.578 / \plot 13.007 18.578 13.045 18.519 / \plot 13.045 18.519 13.081 18.464 / \plot 13.081 18.464 13.113 18.411 / \plot 13.113 18.411 13.142 18.364 / \plot 13.142 18.364 13.170 18.318 / \plot 13.170 18.318 13.195 18.275 / \plot 13.195 18.275 13.219 18.235 / \plot 13.219 18.235 13.242 18.197 / \plot 13.242 18.197 13.261 18.161 / \plot 13.261 18.161 13.282 18.123 / \plot 13.282 18.123 13.303 18.085 / \plot 13.303 18.085 13.324 18.045 / \plot 13.324 18.045 13.346 18.004 / \plot 13.346 18.004 13.365 17.962 / \plot 13.365 17.962 13.386 17.920 / \plot 13.386 17.920 13.405 17.875 / \plot 13.405 17.875 13.424 17.831 / \plot 13.424 17.831 13.443 17.786 / \plot 13.443 17.786 13.462 17.740 / \plot 13.462 17.740 13.479 17.693 / \plot 13.479 17.693 13.496 17.647 / \plot 13.496 17.647 13.511 17.598 / \plot 13.511 17.598 13.526 17.551 / \plot 13.526 17.551 13.540 17.505 / \plot 13.540 17.505 13.553 17.456 / \plot 13.553 17.456 13.564 17.410 / \plot 13.564 17.410 13.574 17.363 / \plot 13.574 17.363 13.583 17.319 / \plot 13.583 17.319 13.591 17.272 / \plot 13.591 17.272 13.600 17.223 / \plot 13.600 17.223 13.606 17.181 / \plot 13.606 17.181 13.612 17.137 / \plot 13.612 17.137 13.617 17.090 / \plot 13.617 17.090 13.621 17.041 / \plot 13.621 17.041 13.625 16.993 / \plot 13.625 16.993 13.629 16.940 / \plot 13.629 16.940 13.631 16.885 / \plot 13.631 16.885 13.636 16.830 / \putrule from 13.636 16.830 to 13.636 16.772 \plot 13.636 16.772 13.638 16.713 / \putrule from 13.638 16.713 to 13.638 16.654 \putrule from 13.638 16.654 to 13.638 16.595 \plot 13.638 16.595 13.636 16.533 / \putrule from 13.636 16.533 to 13.636 16.472 \plot 13.636 16.472 13.631 16.413 / \plot 13.631 16.413 13.629 16.351 / \plot 13.629 16.351 13.625 16.292 / \plot 13.625 16.292 13.621 16.235 / \plot 13.621 16.235 13.617 16.178 / \plot 13.617 16.178 13.612 16.121 / \plot 13.612 16.121 13.606 16.063 / \plot 13.606 16.063 13.600 16.006 / \plot 13.600 16.006 13.593 15.955 / \plot 13.593 15.955 13.587 15.903 / \plot 13.587 15.903 13.578 15.850 / \plot 13.578 15.850 13.570 15.797 / \plot 13.570 15.797 13.561 15.742 / \plot 13.561 15.742 13.553 15.684 / \plot 13.553 15.684 13.542 15.627 / \plot 13.542 15.627 13.530 15.570 / \plot 13.530 15.570 13.519 15.511 / \plot 13.519 15.511 13.506 15.452 / \plot 13.506 15.452 13.494 15.392 / \plot 13.494 15.392 13.479 15.333 / \plot 13.479 15.333 13.464 15.276 / \plot 13.464 15.276 13.449 15.219 / \plot 13.449 15.219 13.434 15.162 / \plot 13.434 15.162 13.418 15.109 / \plot 13.418 15.109 13.403 15.056 / \plot 13.403 15.056 13.386 15.003 / \plot 13.386 15.003 13.369 14.954 / \plot 13.369 14.954 13.352 14.908 / \plot 13.352 14.908 13.335 14.863 / \plot 13.335 14.863 13.318 14.819 / \plot 13.318 14.819 13.299 14.776 / \plot 13.299 14.776 13.282 14.736 / \plot 13.282 14.736 13.259 14.692 / \plot 13.259 14.692 13.238 14.645 / \plot 13.238 14.645 13.212 14.601 / \plot 13.212 14.601 13.185 14.558 / \plot 13.185 14.558 13.157 14.514 / \plot 13.157 14.514 13.125 14.470 / \plot 13.125 14.470 13.089 14.421 / \plot 13.089 14.421 13.051 14.372 / \plot 13.051 14.372 13.011 14.321 / \plot 13.011 14.321 12.967 14.271 / \plot 12.967 14.271 12.922 14.216 / \plot 12.922 14.216 12.876 14.165 / \plot 12.876 14.165 12.833 14.116 / \plot 12.833 14.116 12.793 14.072 / \plot 12.793 14.072 12.759 14.034 / \plot 12.759 14.034 12.734 14.006 / \plot 12.734 14.006 12.715 13.985 / \plot 12.715 13.985 12.704 13.974 / \plot 12.704 13.974 12.700 13.970 / }% \linethickness=1pt \setplotsymbol ({\makebox(0,0)[l]{\tencirc\symbol{'160}}}) {\color[rgb]{0,0,0}\plot 13.652 17.145 13.648 17.151 / \plot 13.648 17.151 13.642 17.164 / \plot 13.642 17.164 13.627 17.189 / \plot 13.627 17.189 13.606 17.225 / \plot 13.606 17.225 13.578 17.274 / \plot 13.578 17.274 13.545 17.333 / \plot 13.545 17.333 13.506 17.399 / \plot 13.506 17.399 13.464 17.471 / \plot 13.464 17.471 13.422 17.543 / \plot 13.422 17.543 13.379 17.615 / \plot 13.379 17.615 13.339 17.685 / \plot 13.339 17.685 13.301 17.752 / \plot 13.301 17.752 13.263 17.814 / \plot 13.263 17.814 13.229 17.871 / \plot 13.229 17.871 13.200 17.926 / \plot 13.200 17.926 13.170 17.975 / \plot 13.170 17.975 13.142 18.019 / \plot 13.142 18.019 13.115 18.062 / \plot 13.115 18.062 13.092 18.102 / \plot 13.092 18.102 13.068 18.140 / \plot 13.068 18.140 13.045 18.176 / \plot 13.045 18.176 13.015 18.220 / \plot 13.015 18.220 12.988 18.265 / \plot 12.988 18.265 12.960 18.307 / \plot 12.960 18.307 12.931 18.349 / \plot 12.931 18.349 12.901 18.392 / \plot 12.901 18.392 12.871 18.432 / \plot 12.871 18.432 12.842 18.472 / \plot 12.842 18.472 12.812 18.512 / \plot 12.812 18.512 12.783 18.553 / \plot 12.783 18.553 12.753 18.589 / \plot 12.753 18.589 12.725 18.625 / \plot 12.725 18.625 12.696 18.658 / \plot 12.696 18.658 12.668 18.688 / \plot 12.668 18.688 12.641 18.718 / \plot 12.641 18.718 12.615 18.743 / \plot 12.615 18.743 12.590 18.768 / \plot 12.590 18.768 12.565 18.792 / \plot 12.565 18.792 12.541 18.811 / \plot 12.541 18.811 12.509 18.836 / \plot 12.509 18.836 12.478 18.860 / \plot 12.478 18.860 12.444 18.881 / \plot 12.444 18.881 12.408 18.902 / \plot 12.408 18.902 12.366 18.923 / \plot 12.366 18.923 12.321 18.946 / \plot 12.321 18.946 12.272 18.967 / \plot 12.272 18.967 12.222 18.989 / \plot 12.222 18.989 12.173 19.010 / \plot 12.173 19.010 12.129 19.027 / \plot 12.129 19.027 12.095 19.039 / \plot 12.095 19.039 12.076 19.046 / \plot 12.076 19.046 12.067 19.050 / \putrule from 12.067 19.050 to 12.065 19.050 }% \put{\SetFigFont{6}{7.2}{\rmdefault}{\mddefault}{\updefault}{\color[rgb]{0,0,0}${\mathcal K}_1$}% } [lB] at 10.058 15.977 \put{\SetFigFont{6}{7.2}{\rmdefault}{\mddefault}{\updefault}{\color[rgb]{0,0,0}$S_+$}% } [lB] at 13.678 15.426 \put{\SetFigFont{6}{7.2}{\rmdefault}{\mddefault}{\updefault}{\color[rgb]{0,0,0}$N_+$}% } [lB] at 13.741 17.141 \put{\SetFigFont{6}{7.2}{\rmdefault}{\mddefault}{\updefault}{\color[rgb]{0,0,0}$S_-$}% } [lB] at 3.243 18.051 \put{\SetFigFont{6}{7.2}{\rmdefault}{\mddefault}{\updefault}{\color[rgb]{0,0,0}$N_-$}% } [lB] at 5.698 18.517 \linethickness=0pt \putrectangle corners at 3.211 19.120 and 13.773 13.877 \endpicture} \qquad \font\thinlinefont=cmr5 \begingroup\makeatletter\ifx\SetFigFont\undefined% \gdef\SetFigFont#1#2#3#4#5{% \reset@font\fontsize{#1}{#2pt}% \fontfamily{#3}\fontseries{#4}\fontshape{#5}% \selectfont}% \fi\endgroup% \mbox{\beginpicture \setcoordinatesystem units <0.50000cm,0.50000cm> \unitlength=0.50000cm \linethickness=1pt \setplotsymbol ({\makebox(0,0)[l]{\tencirc\symbol{'160}}}) \setshadesymbol ({\thinlinefont .}) \setlinear \linethickness= 0.500pt \setplotsymbol ({\thinlinefont .}) {\color[rgb]{0,0,0}\putrule from 5.080 16.510 to 15.240 16.510 }% \linethickness= 0.500pt \setplotsymbol ({\thinlinefont .}) {\color[rgb]{0,0,0}\putrule from 5.080 11.430 to 15.240 11.430 }% \linethickness=1pt \setplotsymbol ({\makebox(0,0)[l]{\tencirc\symbol{'160}}}) {\color[rgb]{0,0,0}\putrule from 5.080 21.590 to 15.240 21.590 }% \linethickness=1pt \setplotsymbol ({\makebox(0,0)[l]{\tencirc\symbol{'160}}}) {\color[rgb]{0,0,0}\putrule from 15.240 21.590 to 15.240 6.315 \putrule from 15.240 6.350 to 5.045 6.350 \putrule from 5.080 6.350 to 5.080 21.590 }% \linethickness= 0.500pt \setplotsymbol ({\thinlinefont .}) {\color[rgb]{0,0,0}\plot 6.985 15.081 7.938 15.716 / \plot 7.761 15.523 7.938 15.716 7.691 15.628 / }% \linethickness= 0.500pt \setplotsymbol ({\thinlinefont .}) {\color[rgb]{0,0,0}\plot 13.970 16.192 14.446 15.875 / \plot 14.200 15.963 14.446 15.875 14.270 16.069 / }% \linethickness= 0.500pt \setplotsymbol ({\thinlinefont .}) {\color[rgb]{0,0,0}\putrule from 5.080 17.462 to 5.080 16.034 \plot 5.017 16.288 5.080 16.034 5.143 16.288 / }% \linethickness= 0.500pt \setplotsymbol ({\thinlinefont .}) {\color[rgb]{0,0,0}\putrule from 15.240 12.383 to 15.240 10.795 \plot 15.176 11.049 15.240 10.795 15.304 11.049 / }% \linethickness= 0.500pt \setplotsymbol ({\thinlinefont .}) {\color[rgb]{0,0,0}\plot 5.080 12.700 5.082 12.706 / \plot 5.082 12.706 5.088 12.717 / \plot 5.088 12.717 5.101 12.740 / \plot 5.101 12.740 5.118 12.774 / \plot 5.118 12.774 5.141 12.819 / \plot 5.141 12.819 5.171 12.876 / \plot 5.171 12.876 5.207 12.941 / \plot 5.207 12.941 5.245 13.013 / \plot 5.245 13.013 5.285 13.089 / \plot 5.285 13.089 5.328 13.168 / \plot 5.328 13.168 5.370 13.244 / \plot 5.370 13.244 5.412 13.320 / \plot 5.412 13.320 5.453 13.392 / \plot 5.453 13.392 5.491 13.460 / \plot 5.491 13.460 5.529 13.523 / \plot 5.529 13.523 5.565 13.583 / \plot 5.565 13.583 5.599 13.638 / \plot 5.599 13.638 5.632 13.691 / \plot 5.632 13.691 5.664 13.739 / \plot 5.664 13.739 5.696 13.786 / \plot 5.696 13.786 5.730 13.830 / \plot 5.730 13.830 5.762 13.875 / \plot 5.762 13.875 5.795 13.917 / \plot 5.795 13.917 5.825 13.955 / \plot 5.825 13.955 5.857 13.995 / \plot 5.857 13.995 5.891 14.034 / \plot 5.891 14.034 5.925 14.074 / \plot 5.925 14.074 5.961 14.112 / \plot 5.961 14.112 5.997 14.152 / \plot 5.997 14.152 6.035 14.192 / \plot 6.035 14.192 6.073 14.232 / \plot 6.073 14.232 6.113 14.275 / \plot 6.113 14.275 6.155 14.315 / \plot 6.155 14.315 6.198 14.357 / \plot 6.198 14.357 6.242 14.398 / \plot 6.242 14.398 6.284 14.440 / \plot 6.284 14.440 6.329 14.480 / \plot 6.329 14.480 6.375 14.520 / \plot 6.375 14.520 6.420 14.561 / \plot 6.420 14.561 6.464 14.601 / \plot 6.464 14.601 6.509 14.639 / \plot 6.509 14.639 6.553 14.677 / \plot 6.553 14.677 6.598 14.713 / \plot 6.598 14.713 6.642 14.751 / \plot 6.642 14.751 6.684 14.785 / \plot 6.684 14.785 6.727 14.821 / \plot 6.727 14.821 6.769 14.855 / \plot 6.769 14.855 6.811 14.889 / \plot 6.811 14.889 6.854 14.922 / \plot 6.854 14.922 6.892 14.954 / \plot 6.892 14.954 6.932 14.986 / \plot 6.932 14.986 6.972 15.018 / \plot 6.972 15.018 7.013 15.050 / \plot 7.013 15.050 7.055 15.081 / \plot 7.055 15.081 7.097 15.115 / \plot 7.097 15.115 7.142 15.149 / \plot 7.142 15.149 7.188 15.183 / \plot 7.188 15.183 7.235 15.219 / \plot 7.235 15.219 7.283 15.255 / \plot 7.283 15.255 7.332 15.291 / \plot 7.332 15.291 7.383 15.329 / \plot 7.383 15.329 7.434 15.365 / \plot 7.434 15.365 7.487 15.403 / \plot 7.487 15.403 7.540 15.439 / \plot 7.540 15.439 7.595 15.477 / \plot 7.595 15.477 7.648 15.515 / \plot 7.648 15.515 7.703 15.551 / \plot 7.703 15.551 7.758 15.589 / \plot 7.758 15.589 7.813 15.625 / \plot 7.813 15.625 7.868 15.661 / \plot 7.868 15.661 7.923 15.697 / \plot 7.923 15.697 7.978 15.731 / \plot 7.978 15.731 8.033 15.765 / \plot 8.033 15.765 8.088 15.801 / \plot 8.088 15.801 8.143 15.835 / \plot 8.143 15.835 8.198 15.869 / \plot 8.198 15.869 8.255 15.900 / \plot 8.255 15.900 8.306 15.932 / \plot 8.306 15.932 8.357 15.962 / \plot 8.357 15.962 8.410 15.991 / \plot 8.410 15.991 8.465 16.021 / \plot 8.465 16.021 8.520 16.053 / \plot 8.520 16.053 8.577 16.085 / \plot 8.577 16.085 8.636 16.116 / \plot 8.636 16.116 8.697 16.148 / \plot 8.697 16.148 8.759 16.180 / \plot 8.759 16.180 8.822 16.212 / \plot 8.822 16.212 8.888 16.245 / \plot 8.888 16.245 8.956 16.277 / \plot 8.956 16.277 9.023 16.309 / \plot 9.023 16.309 9.091 16.341 / \plot 9.091 16.341 9.163 16.375 / \plot 9.163 16.375 9.233 16.404 / \plot 9.233 16.404 9.305 16.436 / \plot 9.305 16.436 9.377 16.466 / \plot 9.377 16.466 9.449 16.495 / \plot 9.449 16.495 9.521 16.525 / \plot 9.521 16.525 9.593 16.552 / \plot 9.593 16.552 9.665 16.578 / \plot 9.665 16.578 9.735 16.603 / \plot 9.735 16.603 9.807 16.626 / \plot 9.807 16.626 9.874 16.650 / \plot 9.874 16.650 9.944 16.671 / \plot 9.944 16.671 10.012 16.692 / \plot 10.012 16.692 10.080 16.711 / \plot 10.080 16.711 10.147 16.728 / \plot 10.147 16.728 10.213 16.745 / \plot 10.213 16.745 10.279 16.760 / \plot 10.279 16.760 10.346 16.775 / \plot 10.346 16.775 10.412 16.787 / \plot 10.412 16.787 10.478 16.800 / \plot 10.478 16.800 10.545 16.811 / \plot 10.545 16.811 10.613 16.821 / \plot 10.613 16.821 10.681 16.830 / \plot 10.681 16.830 10.751 16.836 / \plot 10.751 16.836 10.823 16.842 / \plot 10.823 16.842 10.894 16.849 / \plot 10.894 16.849 10.966 16.851 / \plot 10.966 16.851 11.041 16.855 / \putrule from 11.041 16.855 to 11.115 16.855 \putrule from 11.115 16.855 to 11.191 16.855 \plot 11.191 16.855 11.267 16.853 / \plot 11.267 16.853 11.343 16.849 / \plot 11.343 16.849 11.419 16.844 / \plot 11.419 16.844 11.496 16.838 / \plot 11.496 16.838 11.570 16.830 / \plot 11.570 16.830 11.646 16.819 / \plot 11.646 16.819 11.720 16.808 / \plot 11.720 16.808 11.794 16.796 / \plot 11.794 16.796 11.866 16.781 / \plot 11.866 16.781 11.938 16.766 / \plot 11.938 16.766 12.008 16.749 / \plot 12.008 16.749 12.076 16.730 / \plot 12.076 16.730 12.141 16.711 / \plot 12.141 16.711 12.207 16.690 / \plot 12.207 16.690 12.270 16.667 / \plot 12.270 16.667 12.332 16.643 / \plot 12.332 16.643 12.393 16.618 / \plot 12.393 16.618 12.452 16.593 / \plot 12.452 16.593 12.509 16.565 / \plot 12.509 16.565 12.569 16.535 / \plot 12.569 16.535 12.624 16.506 / \plot 12.624 16.506 12.681 16.474 / \plot 12.681 16.474 12.736 16.440 / \plot 12.736 16.440 12.791 16.404 / \plot 12.791 16.404 12.848 16.366 / \plot 12.848 16.366 12.903 16.328 / \plot 12.903 16.328 12.958 16.286 / \plot 12.958 16.286 13.013 16.241 / \plot 13.013 16.241 13.070 16.195 / \plot 13.070 16.195 13.125 16.148 / \plot 13.125 16.148 13.180 16.097 / \plot 13.180 16.097 13.236 16.044 / \plot 13.236 16.044 13.291 15.991 / \plot 13.291 15.991 13.343 15.936 / \plot 13.343 15.936 13.396 15.879 / \plot 13.396 15.879 13.449 15.820 / \plot 13.449 15.820 13.502 15.761 / \plot 13.502 15.761 13.551 15.699 / \plot 13.551 15.699 13.602 15.638 / \plot 13.602 15.638 13.648 15.574 / \plot 13.648 15.574 13.695 15.513 / \plot 13.695 15.513 13.739 15.450 / \plot 13.739 15.450 13.784 15.386 / \plot 13.784 15.386 13.824 15.323 / \plot 13.824 15.323 13.864 15.259 / \plot 13.864 15.259 13.902 15.196 / \plot 13.902 15.196 13.940 15.132 / \plot 13.940 15.132 13.974 15.069 / \plot 13.974 15.069 14.008 15.007 / \plot 14.008 15.007 14.042 14.944 / \plot 14.042 14.944 14.072 14.880 / \plot 14.072 14.880 14.103 14.817 / \plot 14.103 14.817 14.129 14.760 / \plot 14.129 14.760 14.154 14.702 / \plot 14.154 14.702 14.177 14.643 / \plot 14.177 14.643 14.203 14.584 / \plot 14.203 14.584 14.226 14.522 / \plot 14.226 14.522 14.249 14.461 / \plot 14.249 14.461 14.273 14.398 / \plot 14.273 14.398 14.294 14.332 / \plot 14.294 14.332 14.317 14.264 / \plot 14.317 14.264 14.338 14.196 / \plot 14.338 14.196 14.359 14.127 / \plot 14.359 14.127 14.381 14.055 / \plot 14.381 14.055 14.402 13.981 / \plot 14.402 13.981 14.423 13.904 / \plot 14.423 13.904 14.442 13.826 / \plot 14.442 13.826 14.461 13.748 / \plot 14.461 13.748 14.482 13.667 / \plot 14.482 13.667 14.499 13.587 / \plot 14.499 13.587 14.518 13.504 / \plot 14.518 13.504 14.535 13.420 / \plot 14.535 13.420 14.554 13.335 / \plot 14.554 13.335 14.571 13.250 / \plot 14.571 13.250 14.586 13.164 / \plot 14.586 13.164 14.603 13.079 / \plot 14.603 13.079 14.618 12.992 / \plot 14.618 12.992 14.633 12.903 / \plot 14.633 12.903 14.647 12.816 / \plot 14.647 12.816 14.662 12.730 / \plot 14.662 12.730 14.675 12.641 / \plot 14.675 12.641 14.690 12.552 / \plot 14.690 12.552 14.702 12.463 / \plot 14.702 12.463 14.715 12.374 / \plot 14.715 12.374 14.728 12.285 / \plot 14.728 12.285 14.738 12.196 / \plot 14.738 12.196 14.751 12.105 / \plot 14.751 12.105 14.764 12.012 / \plot 14.764 12.012 14.774 11.936 / \plot 14.774 11.936 14.783 11.858 / \plot 14.783 11.858 14.793 11.779 / \plot 14.793 11.779 14.804 11.697 / \plot 14.804 11.697 14.815 11.614 / \plot 14.815 11.614 14.825 11.529 / \plot 14.825 11.529 14.834 11.441 / \plot 14.834 11.441 14.846 11.350 / \plot 14.846 11.350 14.857 11.256 / \plot 14.857 11.256 14.867 11.157 / \plot 14.867 11.157 14.878 11.055 / \plot 14.878 11.055 14.891 10.950 / \plot 14.891 10.950 14.903 10.839 / \plot 14.903 10.839 14.916 10.723 / \plot 14.916 10.723 14.929 10.604 / \plot 14.929 10.604 14.942 10.480 / \plot 14.942 10.480 14.956 10.348 / \plot 14.956 10.348 14.971 10.215 / \plot 14.971 10.215 14.986 10.075 / \plot 14.986 10.075 15.001 9.931 / \plot 15.001 9.931 15.016 9.783 / \plot 15.016 9.783 15.033 9.631 / \plot 15.033 9.631 15.050 9.478 / \plot 15.050 9.478 15.064 9.322 / \plot 15.064 9.322 15.081 9.165 / \plot 15.081 9.165 15.098 9.011 / \plot 15.098 9.011 15.113 8.856 / \plot 15.113 8.856 15.128 8.706 / \plot 15.128 8.706 15.143 8.560 / \plot 15.143 8.560 15.157 8.422 / \plot 15.157 8.422 15.172 8.293 / \plot 15.172 8.293 15.183 8.172 / \plot 15.183 8.172 15.196 8.064 / \plot 15.196 8.064 15.204 7.967 / \plot 15.204 7.967 15.212 7.882 / \plot 15.212 7.882 15.221 7.811 / \plot 15.221 7.811 15.227 7.751 / \plot 15.227 7.751 15.232 7.705 / \plot 15.232 7.705 15.236 7.671 / \plot 15.236 7.671 15.238 7.648 / \putrule from 15.238 7.648 to 15.238 7.631 \plot 15.238 7.631 15.240 7.624 / \putrule from 15.240 7.624 to 15.240 7.620 }% \linethickness= 0.500pt \setplotsymbol ({\thinlinefont .}) {\color[rgb]{0,0,0}\plot 15.240 15.240 15.236 15.242 / \plot 15.236 15.242 15.229 15.248 / \plot 15.229 15.248 15.215 15.261 / \plot 15.215 15.261 15.193 15.278 / \plot 15.193 15.278 15.164 15.304 / \plot 15.164 15.304 15.124 15.335 / \plot 15.124 15.335 15.077 15.373 / \plot 15.077 15.373 15.024 15.418 / \plot 15.024 15.418 14.963 15.466 / \plot 14.963 15.466 14.897 15.519 / \plot 14.897 15.519 14.829 15.574 / \plot 14.829 15.574 14.760 15.632 / \plot 14.760 15.632 14.688 15.687 / \plot 14.688 15.687 14.616 15.744 / \plot 14.616 15.744 14.546 15.797 / \plot 14.546 15.797 14.476 15.852 / \plot 14.476 15.852 14.408 15.903 / \plot 14.408 15.903 14.343 15.953 / \plot 14.343 15.953 14.275 16.002 / \plot 14.275 16.002 14.211 16.049 / \plot 14.211 16.049 14.146 16.095 / \plot 14.146 16.095 14.080 16.142 / \plot 14.080 16.142 14.014 16.188 / \plot 14.014 16.188 13.947 16.235 / \plot 13.947 16.235 13.877 16.281 / \plot 13.877 16.281 13.805 16.328 / \plot 13.805 16.328 13.733 16.377 / \plot 13.733 16.377 13.678 16.413 / \plot 13.678 16.413 13.623 16.449 / \plot 13.623 16.449 13.568 16.485 / \plot 13.568 16.485 13.509 16.521 / \plot 13.509 16.521 13.447 16.559 / \plot 13.447 16.559 13.386 16.597 / \plot 13.386 16.597 13.322 16.635 / \plot 13.322 16.635 13.255 16.673 / \plot 13.255 16.673 13.187 16.711 / \plot 13.187 16.711 13.115 16.749 / \plot 13.115 16.749 13.041 16.789 / \plot 13.041 16.789 12.967 16.828 / \plot 12.967 16.828 12.888 16.866 / \plot 12.888 16.866 12.808 16.904 / \plot 12.808 16.904 12.725 16.942 / \plot 12.725 16.942 12.641 16.978 / \plot 12.641 16.978 12.556 17.014 / \plot 12.556 17.014 12.467 17.050 / \plot 12.467 17.050 12.378 17.081 / \plot 12.378 17.081 12.287 17.115 / \plot 12.287 17.115 12.194 17.145 / \plot 12.194 17.145 12.101 17.175 / \plot 12.101 17.175 12.006 17.200 / \plot 12.006 17.200 11.910 17.225 / \plot 11.910 17.225 11.815 17.249 / \plot 11.815 17.249 11.718 17.268 / \plot 11.718 17.268 11.620 17.287 / \plot 11.620 17.287 11.523 17.304 / \plot 11.523 17.304 11.426 17.316 / \plot 11.426 17.316 11.328 17.327 / \plot 11.328 17.327 11.229 17.333 / \plot 11.229 17.333 11.132 17.340 / \plot 11.132 17.340 11.034 17.342 / \putrule from 11.034 17.342 to 10.937 17.342 \plot 10.937 17.342 10.837 17.338 / \plot 10.837 17.338 10.740 17.331 / \plot 10.740 17.331 10.643 17.323 / \plot 10.643 17.323 10.543 17.310 / \plot 10.543 17.310 10.446 17.295 / \plot 10.446 17.295 10.346 17.276 / \plot 10.346 17.276 10.253 17.257 / \plot 10.253 17.257 10.162 17.236 / \plot 10.162 17.236 10.069 17.211 / \plot 10.069 17.211 9.974 17.183 / \plot 9.974 17.183 9.878 17.156 / \plot 9.878 17.156 9.781 17.124 / \plot 9.781 17.124 9.682 17.090 / \plot 9.682 17.090 9.582 17.054 / \plot 9.582 17.054 9.481 17.016 / \plot 9.481 17.016 9.379 16.978 / \plot 9.379 16.978 9.273 16.938 / \plot 9.273 16.938 9.169 16.895 / \plot 9.169 16.895 9.064 16.851 / \plot 9.064 16.851 8.956 16.806 / \plot 8.956 16.806 8.850 16.762 / \plot 8.850 16.762 8.740 16.717 / \plot 8.740 16.717 8.632 16.671 / \plot 8.632 16.671 8.524 16.626 / \plot 8.524 16.626 8.414 16.580 / \plot 8.414 16.580 8.306 16.535 / \plot 8.306 16.535 8.198 16.491 / \plot 8.198 16.491 8.090 16.447 / \plot 8.090 16.447 7.984 16.404 / \plot 7.984 16.404 7.878 16.362 / \plot 7.878 16.362 7.775 16.324 / \plot 7.775 16.324 7.671 16.286 / \plot 7.671 16.286 7.569 16.250 / \plot 7.569 16.250 7.470 16.216 / \plot 7.470 16.216 7.374 16.184 / \plot 7.374 16.184 7.279 16.157 / \plot 7.279 16.157 7.186 16.131 / \plot 7.186 16.131 7.095 16.108 / \plot 7.095 16.108 7.008 16.089 / \plot 7.008 16.089 6.924 16.072 / \plot 6.924 16.072 6.841 16.059 / \plot 6.841 16.059 6.761 16.051 / \plot 6.761 16.051 6.682 16.044 / \plot 6.682 16.044 6.608 16.042 / \plot 6.608 16.042 6.538 16.044 / \plot 6.538 16.044 6.469 16.049 / \plot 6.469 16.049 6.403 16.059 / \plot 6.403 16.059 6.337 16.072 / \plot 6.337 16.072 6.276 16.091 / \plot 6.276 16.091 6.219 16.112 / \plot 6.219 16.112 6.164 16.140 / \plot 6.164 16.140 6.111 16.169 / \plot 6.111 16.169 6.060 16.205 / \plot 6.060 16.205 6.013 16.245 / \plot 6.013 16.245 5.967 16.290 / \plot 5.967 16.290 5.922 16.341 / \plot 5.922 16.341 5.880 16.398 / \plot 5.880 16.398 5.840 16.461 / \plot 5.840 16.461 5.800 16.529 / \plot 5.800 16.529 5.762 16.605 / \plot 5.762 16.605 5.726 16.688 / \plot 5.726 16.688 5.690 16.777 / \plot 5.690 16.777 5.656 16.872 / \plot 5.656 16.872 5.622 16.976 / \plot 5.622 16.976 5.590 17.086 / \plot 5.590 17.086 5.558 17.204 / \plot 5.558 17.204 5.529 17.329 / \plot 5.529 17.329 5.499 17.462 / \plot 5.499 17.462 5.469 17.602 / \plot 5.469 17.602 5.442 17.748 / \plot 5.442 17.748 5.412 17.903 / \plot 5.412 17.903 5.387 18.062 / \plot 5.387 18.062 5.359 18.227 / \plot 5.359 18.227 5.334 18.396 / \plot 5.334 18.396 5.309 18.570 / \plot 5.309 18.570 5.285 18.743 / \plot 5.285 18.743 5.262 18.919 / \plot 5.262 18.919 5.241 19.094 / \plot 5.241 19.094 5.220 19.266 / \plot 5.220 19.266 5.201 19.435 / \plot 5.201 19.435 5.182 19.596 / \plot 5.182 19.596 5.165 19.751 / \plot 5.165 19.751 5.150 19.895 / \plot 5.150 19.895 5.137 20.028 / \plot 5.137 20.028 5.124 20.149 / \plot 5.124 20.149 5.114 20.256 / \plot 5.114 20.256 5.105 20.350 / \plot 5.105 20.350 5.099 20.428 / \plot 5.099 20.428 5.093 20.494 / \plot 5.093 20.494 5.088 20.544 / \plot 5.088 20.544 5.084 20.582 / \plot 5.084 20.582 5.082 20.608 / \plot 5.082 20.608 5.080 20.625 / \putrule from 5.080 20.625 to 5.080 20.633 \putrule from 5.080 20.633 to 5.080 20.637 }% \linethickness= 0.500pt \setplotsymbol ({\thinlinefont .}) {\color[rgb]{0,0,0}\putrule from 15.240 17.780 to 15.240 17.784 \putrule from 15.240 17.784 to 15.240 17.795 \plot 15.240 17.795 15.238 17.814 / \plot 15.238 17.814 15.236 17.844 / \plot 15.236 17.844 15.234 17.884 / \plot 15.234 17.884 15.229 17.937 / \plot 15.229 17.937 15.225 18.002 / \plot 15.225 18.002 15.219 18.078 / \plot 15.219 18.078 15.215 18.165 / \plot 15.215 18.165 15.208 18.260 / \plot 15.208 18.260 15.200 18.362 / \plot 15.200 18.362 15.193 18.468 / \plot 15.193 18.468 15.185 18.576 / \plot 15.185 18.576 15.179 18.684 / \plot 15.179 18.684 15.170 18.790 / \plot 15.170 18.790 15.164 18.895 / \plot 15.164 18.895 15.157 18.995 / \plot 15.157 18.995 15.149 19.092 / \plot 15.149 19.092 15.143 19.185 / \plot 15.143 19.185 15.136 19.274 / \plot 15.136 19.274 15.130 19.357 / \plot 15.130 19.357 15.126 19.435 / \plot 15.126 19.435 15.119 19.509 / \plot 15.119 19.509 15.115 19.581 / \plot 15.115 19.581 15.109 19.649 / \plot 15.109 19.649 15.105 19.713 / \plot 15.105 19.713 15.100 19.774 / \plot 15.100 19.774 15.094 19.833 / \plot 15.094 19.833 15.090 19.890 / \plot 15.090 19.890 15.085 19.947 / \plot 15.085 19.947 15.081 20.003 / \plot 15.081 20.003 15.075 20.068 / \plot 15.075 20.068 15.071 20.134 / \plot 15.071 20.134 15.064 20.199 / \plot 15.064 20.199 15.058 20.265 / \plot 15.058 20.265 15.052 20.331 / \plot 15.052 20.331 15.045 20.398 / \plot 15.045 20.398 15.039 20.466 / \plot 15.039 20.466 15.033 20.538 / \plot 15.033 20.538 15.026 20.612 / \plot 15.026 20.612 15.018 20.690 / \plot 15.018 20.690 15.009 20.771 / \plot 15.009 20.771 15.001 20.853 / \plot 15.001 20.853 14.992 20.940 / \plot 14.992 20.940 14.982 21.027 / \plot 14.982 21.027 14.973 21.114 / \plot 14.973 21.114 14.965 21.201 / \plot 14.965 21.201 14.956 21.281 / \plot 14.956 21.281 14.948 21.355 / \plot 14.948 21.355 14.942 21.421 / \plot 14.942 21.421 14.935 21.476 / \plot 14.935 21.476 14.931 21.520 / \plot 14.931 21.520 14.927 21.552 / \plot 14.927 21.552 14.925 21.573 / \plot 14.925 21.573 14.922 21.586 / \putrule from 14.922 21.586 to 14.922 21.590 }% \linethickness= 0.500pt \setplotsymbol ({\thinlinefont .}) {\color[rgb]{0,0,0}\putrule from 5.080 10.160 to 5.080 10.156 \plot 5.080 10.156 5.082 10.145 / \plot 5.082 10.145 5.084 10.126 / \plot 5.084 10.126 5.086 10.096 / \plot 5.086 10.096 5.093 10.056 / \plot 5.093 10.056 5.099 10.001 / \plot 5.099 10.001 5.105 9.934 / \plot 5.105 9.934 5.114 9.855 / \plot 5.114 9.855 5.124 9.764 / \plot 5.124 9.764 5.137 9.665 / \plot 5.137 9.665 5.150 9.557 / \plot 5.150 9.557 5.163 9.442 / \plot 5.163 9.442 5.177 9.326 / \plot 5.177 9.326 5.190 9.205 / \plot 5.190 9.205 5.205 9.087 / \plot 5.205 9.087 5.218 8.968 / \plot 5.218 8.968 5.232 8.854 / \plot 5.232 8.854 5.245 8.744 / \plot 5.245 8.744 5.258 8.638 / \plot 5.258 8.638 5.271 8.537 / \plot 5.271 8.537 5.283 8.439 / \plot 5.283 8.439 5.294 8.348 / \plot 5.294 8.348 5.304 8.263 / \plot 5.304 8.263 5.315 8.183 / \plot 5.315 8.183 5.326 8.107 / \plot 5.326 8.107 5.336 8.035 / \plot 5.336 8.035 5.345 7.965 / \plot 5.345 7.965 5.353 7.902 / \plot 5.353 7.902 5.364 7.840 / \plot 5.364 7.840 5.372 7.783 / \plot 5.372 7.783 5.381 7.726 / \plot 5.381 7.726 5.389 7.673 / \plot 5.389 7.673 5.397 7.620 / \plot 5.397 7.620 5.410 7.548 / \plot 5.410 7.548 5.423 7.480 / \plot 5.423 7.480 5.433 7.415 / \plot 5.433 7.415 5.446 7.349 / \plot 5.446 7.349 5.461 7.286 / \plot 5.461 7.286 5.474 7.222 / \plot 5.474 7.222 5.489 7.159 / \plot 5.489 7.159 5.505 7.093 / \plot 5.505 7.093 5.520 7.027 / \plot 5.520 7.027 5.539 6.960 / \plot 5.539 6.960 5.558 6.890 / \plot 5.558 6.890 5.577 6.820 / \plot 5.577 6.820 5.596 6.750 / \plot 5.596 6.750 5.616 6.680 / \plot 5.616 6.680 5.637 6.612 / \plot 5.637 6.612 5.654 6.551 / \plot 5.654 6.551 5.671 6.496 / \plot 5.671 6.496 5.685 6.447 / \plot 5.685 6.447 5.696 6.411 / \plot 5.696 6.411 5.704 6.384 / \plot 5.704 6.384 5.711 6.365 / \plot 5.711 6.365 5.713 6.354 / \plot 5.713 6.354 5.715 6.350 / }% \linethickness= 0.500pt \setplotsymbol ({\thinlinefont .}) {\color[rgb]{0,0,0}\plot 8.255 6.350 8.259 6.352 / \plot 8.259 6.352 8.270 6.354 / \plot 8.270 6.354 8.289 6.361 / \plot 8.289 6.361 8.316 6.369 / \plot 8.316 6.369 8.357 6.380 / \plot 8.357 6.380 8.407 6.394 / \plot 8.407 6.394 8.467 6.411 / \plot 8.467 6.411 8.537 6.433 / \plot 8.537 6.433 8.613 6.456 / \plot 8.613 6.456 8.695 6.479 / \plot 8.695 6.479 8.780 6.505 / \plot 8.780 6.505 8.867 6.528 / \plot 8.867 6.528 8.951 6.553 / \plot 8.951 6.553 9.036 6.576 / \plot 9.036 6.576 9.116 6.600 / \plot 9.116 6.600 9.195 6.623 / \plot 9.195 6.623 9.269 6.642 / \plot 9.269 6.642 9.339 6.663 / \plot 9.339 6.663 9.406 6.680 / \plot 9.406 6.680 9.470 6.699 / \plot 9.470 6.699 9.533 6.714 / \plot 9.533 6.714 9.591 6.731 / \plot 9.591 6.731 9.650 6.746 / \plot 9.650 6.746 9.705 6.761 / \plot 9.705 6.761 9.760 6.773 / \plot 9.760 6.773 9.815 6.786 / \plot 9.815 6.786 9.870 6.799 / \plot 9.870 6.799 9.923 6.814 / \plot 9.923 6.814 9.978 6.826 / \plot 9.978 6.826 10.035 6.839 / \plot 10.035 6.839 10.092 6.852 / \plot 10.092 6.852 10.149 6.862 / \plot 10.149 6.862 10.209 6.875 / \plot 10.209 6.875 10.268 6.888 / \plot 10.268 6.888 10.329 6.900 / \plot 10.329 6.900 10.393 6.911 / \plot 10.393 6.911 10.456 6.924 / \plot 10.456 6.924 10.520 6.934 / \plot 10.520 6.934 10.585 6.947 / \plot 10.585 6.947 10.651 6.957 / \plot 10.651 6.957 10.717 6.966 / \plot 10.717 6.966 10.780 6.977 / \plot 10.780 6.977 10.846 6.985 / \plot 10.846 6.985 10.911 6.993 / \plot 10.911 6.993 10.975 7.002 / \plot 10.975 7.002 11.038 7.008 / \plot 11.038 7.008 11.102 7.015 / \plot 11.102 7.015 11.163 7.021 / \plot 11.163 7.021 11.223 7.025 / \plot 11.223 7.025 11.282 7.029 / \plot 11.282 7.029 11.339 7.032 / \plot 11.339 7.032 11.396 7.034 / \plot 11.396 7.034 11.453 7.036 / \plot 11.453 7.036 11.508 7.038 / \putrule from 11.508 7.038 to 11.563 7.038 \putrule from 11.563 7.038 to 11.620 7.038 \plot 11.620 7.038 11.680 7.036 / \plot 11.680 7.036 11.739 7.034 / \plot 11.739 7.034 11.798 7.032 / \plot 11.798 7.032 11.860 7.027 / \plot 11.860 7.027 11.921 7.023 / \plot 11.921 7.023 11.982 7.017 / \plot 11.982 7.017 12.046 7.010 / \plot 12.046 7.010 12.107 7.004 / \plot 12.107 7.004 12.171 6.996 / \plot 12.171 6.996 12.234 6.987 / \plot 12.234 6.987 12.298 6.977 / \plot 12.298 6.977 12.361 6.966 / \plot 12.361 6.966 12.423 6.955 / \plot 12.423 6.955 12.484 6.945 / \plot 12.484 6.945 12.543 6.932 / \plot 12.543 6.932 12.601 6.919 / \plot 12.601 6.919 12.658 6.907 / \plot 12.658 6.907 12.713 6.894 / \plot 12.713 6.894 12.766 6.881 / \plot 12.766 6.881 12.816 6.869 / \plot 12.816 6.869 12.865 6.854 / \plot 12.865 6.854 12.912 6.841 / \plot 12.912 6.841 12.958 6.828 / \plot 12.958 6.828 13.001 6.814 / \plot 13.001 6.814 13.045 6.799 / \plot 13.045 6.799 13.094 6.784 / \plot 13.094 6.784 13.140 6.765 / \plot 13.140 6.765 13.187 6.748 / \plot 13.187 6.748 13.236 6.729 / \plot 13.236 6.729 13.282 6.710 / \plot 13.282 6.710 13.331 6.687 / \plot 13.331 6.687 13.379 6.665 / \plot 13.379 6.665 13.432 6.640 / \plot 13.432 6.640 13.487 6.612 / \plot 13.487 6.612 13.542 6.583 / \plot 13.542 6.583 13.602 6.553 / \plot 13.602 6.553 13.661 6.521 / \plot 13.661 6.521 13.718 6.490 / \plot 13.718 6.490 13.775 6.460 / \plot 13.775 6.460 13.826 6.430 / \plot 13.826 6.430 13.873 6.405 / \plot 13.873 6.405 13.909 6.384 / \plot 13.909 6.384 13.936 6.369 / \plot 13.936 6.369 13.955 6.358 / \plot 13.955 6.358 13.966 6.352 / \plot 13.966 6.352 13.970 6.350 / }% \linethickness= 0.500pt \setplotsymbol ({\thinlinefont .}) {\color[rgb]{0,0,0}\plot 9.525 6.350 9.531 6.352 / \plot 9.531 6.352 9.544 6.354 / \plot 9.544 6.354 9.565 6.361 / \plot 9.565 6.361 9.601 6.367 / \plot 9.601 6.367 9.648 6.378 / \plot 9.648 6.378 9.705 6.392 / \plot 9.705 6.392 9.773 6.407 / \plot 9.773 6.407 9.847 6.422 / \plot 9.847 6.422 9.925 6.441 / \plot 9.925 6.441 10.008 6.458 / \plot 10.008 6.458 10.088 6.475 / \plot 10.088 6.475 10.166 6.492 / \plot 10.166 6.492 10.243 6.507 / \plot 10.243 6.507 10.317 6.519 / \plot 10.317 6.519 10.384 6.532 / \plot 10.384 6.532 10.450 6.543 / \plot 10.450 6.543 10.513 6.553 / \plot 10.513 6.553 10.573 6.562 / \plot 10.573 6.562 10.630 6.568 / \plot 10.630 6.568 10.685 6.574 / \plot 10.685 6.574 10.740 6.581 / \plot 10.740 6.581 10.793 6.585 / \plot 10.793 6.585 10.848 6.587 / \plot 10.848 6.587 10.903 6.591 / \plot 10.903 6.591 10.958 6.593 / \putrule from 10.958 6.593 to 11.015 6.593 \plot 11.015 6.593 11.072 6.596 / \putrule from 11.072 6.596 to 11.134 6.596 \plot 11.134 6.596 11.193 6.593 / \putrule from 11.193 6.593 to 11.254 6.593 \plot 11.254 6.593 11.318 6.591 / \plot 11.318 6.591 11.381 6.587 / \plot 11.381 6.587 11.445 6.585 / \plot 11.445 6.585 11.510 6.581 / \plot 11.510 6.581 11.574 6.576 / \plot 11.574 6.576 11.637 6.572 / \plot 11.637 6.572 11.701 6.568 / \plot 11.701 6.568 11.762 6.562 / \plot 11.762 6.562 11.822 6.557 / \plot 11.822 6.557 11.879 6.551 / \plot 11.879 6.551 11.934 6.545 / \plot 11.934 6.545 11.989 6.538 / \plot 11.989 6.538 12.040 6.532 / \plot 12.040 6.532 12.088 6.528 / \plot 12.088 6.528 12.135 6.521 / \plot 12.135 6.521 12.181 6.515 / \plot 12.181 6.515 12.224 6.509 / \plot 12.224 6.509 12.279 6.500 / \plot 12.279 6.500 12.332 6.492 / \plot 12.332 6.492 12.385 6.483 / \plot 12.385 6.483 12.435 6.475 / \plot 12.435 6.475 12.488 6.466 / \plot 12.488 6.466 12.541 6.456 / \plot 12.541 6.456 12.596 6.443 / \plot 12.596 6.443 12.653 6.433 / \plot 12.653 6.433 12.713 6.420 / \plot 12.713 6.420 12.774 6.405 / \plot 12.774 6.405 12.831 6.392 / \plot 12.831 6.392 12.886 6.380 / \plot 12.886 6.380 12.933 6.369 / \plot 12.933 6.369 12.971 6.361 / \plot 12.971 6.361 12.996 6.354 / \plot 12.996 6.354 13.011 6.352 / \plot 13.011 6.352 13.018 6.350 / }% \linethickness=0pt \setplotsymbol ({\thinlinefont \ }) {\color[rgb]{0,0,0}\plot 5.397 17.939 5.408 17.867 / \plot 5.408 17.867 5.421 17.795 / \plot 5.421 17.795 5.431 17.725 / \plot 5.431 17.725 5.444 17.653 / \plot 5.444 17.653 5.455 17.583 / \plot 5.455 17.583 5.467 17.513 / \plot 5.467 17.513 5.480 17.443 / \plot 5.480 17.443 5.493 17.374 / \plot 5.493 17.374 5.508 17.306 / \plot 5.508 17.306 5.520 17.238 / \plot 5.520 17.238 5.535 17.170 / \plot 5.535 17.170 5.550 17.105 / \plot 5.550 17.105 5.563 17.041 / \plot 5.563 17.041 5.580 16.980 / \plot 5.580 16.980 5.594 16.919 / \plot 5.594 16.919 5.609 16.861 / \plot 5.609 16.861 5.624 16.806 / \plot 5.624 16.806 5.641 16.753 / \plot 5.641 16.753 5.656 16.705 / \plot 5.656 16.705 5.671 16.656 / \plot 5.671 16.656 5.687 16.612 / \plot 5.687 16.612 5.702 16.571 / \plot 5.702 16.571 5.719 16.533 / \plot 5.719 16.533 5.734 16.497 / \plot 5.734 16.497 5.751 16.463 / \plot 5.751 16.463 5.768 16.430 / \plot 5.768 16.430 5.791 16.391 / \plot 5.791 16.391 5.812 16.355 / \plot 5.812 16.355 5.838 16.324 / \plot 5.838 16.324 5.863 16.292 / \plot 5.863 16.292 5.891 16.262 / \plot 5.891 16.262 5.920 16.237 / \plot 5.920 16.237 5.950 16.212 / \plot 5.950 16.212 5.982 16.188 / \plot 5.982 16.188 6.016 16.169 / \plot 6.016 16.169 6.052 16.150 / \plot 6.052 16.150 6.088 16.133 / \plot 6.088 16.133 6.124 16.118 / \plot 6.124 16.118 6.159 16.108 / \plot 6.159 16.108 6.198 16.097 / \plot 6.198 16.097 6.236 16.087 / \plot 6.236 16.087 6.274 16.080 / \plot 6.274 16.080 6.312 16.074 / \plot 6.312 16.074 6.350 16.068 / \plot 6.350 16.068 6.390 16.063 / \plot 6.390 16.063 6.430 16.059 / \plot 6.430 16.059 6.462 16.057 / \plot 6.462 16.057 6.498 16.055 / \plot 6.498 16.055 6.534 16.053 / \plot 6.534 16.053 6.572 16.051 / \plot 6.572 16.051 6.612 16.049 / \putrule from 6.612 16.049 to 6.655 16.049 \putrule from 6.655 16.049 to 6.699 16.049 \plot 6.699 16.049 6.746 16.051 / \putrule from 6.746 16.051 to 6.794 16.051 \plot 6.794 16.051 6.845 16.055 / \plot 6.845 16.055 6.898 16.059 / \plot 6.898 16.059 6.953 16.063 / \plot 6.953 16.063 7.010 16.070 / \plot 7.010 16.070 7.070 16.076 / \plot 7.070 16.076 7.129 16.085 / \plot 7.129 16.085 7.192 16.095 / \plot 7.192 16.095 7.254 16.106 / \plot 7.254 16.106 7.317 16.118 / \plot 7.317 16.118 7.383 16.131 / \plot 7.383 16.131 7.449 16.146 / \plot 7.449 16.146 7.516 16.163 / \plot 7.516 16.163 7.584 16.180 / \plot 7.584 16.180 7.654 16.199 / \plot 7.654 16.199 7.726 16.218 / \plot 7.726 16.218 7.781 16.235 / \plot 7.781 16.235 7.838 16.252 / \plot 7.838 16.252 7.897 16.271 / \plot 7.897 16.271 7.957 16.290 / \plot 7.957 16.290 8.020 16.311 / \plot 8.020 16.311 8.084 16.332 / \plot 8.084 16.332 8.149 16.355 / \plot 8.149 16.355 8.217 16.379 / \plot 8.217 16.379 8.287 16.404 / \plot 8.287 16.404 8.359 16.430 / \plot 8.359 16.430 8.431 16.455 / \plot 8.431 16.455 8.505 16.480 / \plot 8.505 16.480 8.579 16.508 / \plot 8.579 16.508 8.655 16.538 / \plot 8.655 16.538 8.731 16.565 / \plot 8.731 16.565 8.807 16.593 / \plot 8.807 16.593 8.884 16.622 / \plot 8.884 16.622 8.960 16.650 / \plot 8.960 16.650 9.034 16.677 / \plot 9.034 16.677 9.108 16.707 / \plot 9.108 16.707 9.182 16.734 / \plot 9.182 16.734 9.254 16.760 / \plot 9.254 16.760 9.324 16.787 / \plot 9.324 16.787 9.392 16.813 / \plot 9.392 16.813 9.459 16.838 / \plot 9.459 16.838 9.523 16.861 / \plot 9.523 16.861 9.586 16.885 / \plot 9.586 16.885 9.646 16.906 / \plot 9.646 16.906 9.705 16.927 / \plot 9.705 16.927 9.762 16.948 / \plot 9.762 16.948 9.815 16.967 / \plot 9.815 16.967 9.870 16.986 / \plot 9.870 16.986 9.938 17.010 / \plot 9.938 17.010 10.003 17.033 / \plot 10.003 17.033 10.069 17.054 / \plot 10.069 17.054 10.132 17.073 / \plot 10.132 17.073 10.194 17.092 / \plot 10.194 17.092 10.255 17.109 / \plot 10.255 17.109 10.317 17.126 / \plot 10.317 17.126 10.376 17.141 / \plot 10.376 17.141 10.435 17.156 / \plot 10.435 17.156 10.494 17.168 / \plot 10.494 17.168 10.552 17.181 / \plot 10.552 17.181 10.607 17.192 / \plot 10.607 17.192 10.662 17.200 / \plot 10.662 17.200 10.715 17.209 / \plot 10.715 17.209 10.767 17.215 / \plot 10.767 17.215 10.818 17.221 / \plot 10.818 17.221 10.869 17.225 / \plot 10.869 17.225 10.918 17.228 / \plot 10.918 17.228 10.964 17.230 / \putrule from 10.964 17.230 to 11.011 17.230 \putrule from 11.011 17.230 to 11.057 17.230 \putrule from 11.057 17.230 to 11.102 17.230 \plot 11.102 17.230 11.146 17.228 / \plot 11.146 17.228 11.193 17.223 / \plot 11.193 17.223 11.235 17.221 / \plot 11.235 17.221 11.278 17.217 / \plot 11.278 17.217 11.322 17.213 / \plot 11.322 17.213 11.369 17.209 / \plot 11.369 17.209 11.415 17.202 / \plot 11.415 17.202 11.464 17.194 / \plot 11.464 17.194 11.515 17.185 / \plot 11.515 17.185 11.568 17.177 / \plot 11.568 17.177 11.620 17.166 / \plot 11.620 17.166 11.676 17.156 / \plot 11.676 17.156 11.733 17.143 / \plot 11.733 17.143 11.790 17.130 / \plot 11.790 17.130 11.849 17.115 / \plot 11.849 17.115 11.908 17.101 / \plot 11.908 17.101 11.968 17.084 / \plot 11.968 17.084 12.029 17.067 / \plot 12.029 17.067 12.088 17.050 / \plot 12.088 17.050 12.150 17.031 / \plot 12.150 17.031 12.209 17.010 / \plot 12.209 17.010 12.268 16.990 / \plot 12.268 16.990 12.327 16.969 / \plot 12.327 16.969 12.387 16.948 / \plot 12.387 16.948 12.446 16.925 / \plot 12.446 16.925 12.503 16.902 / \plot 12.503 16.902 12.562 16.878 / \plot 12.562 16.878 12.622 16.853 / \plot 12.622 16.853 12.675 16.830 / \plot 12.675 16.830 12.732 16.804 / \plot 12.732 16.804 12.789 16.779 / \plot 12.789 16.779 12.846 16.751 / \plot 12.846 16.751 12.905 16.724 / \plot 12.905 16.724 12.965 16.694 / \plot 12.965 16.694 13.026 16.662 / \plot 13.026 16.662 13.089 16.631 / \plot 13.089 16.631 13.153 16.599 / \plot 13.153 16.599 13.216 16.565 / \plot 13.216 16.565 13.282 16.529 / \plot 13.282 16.529 13.348 16.493 / \plot 13.348 16.493 13.413 16.457 / \plot 13.413 16.457 13.479 16.419 / \plot 13.479 16.419 13.545 16.383 / \plot 13.545 16.383 13.608 16.345 / \plot 13.608 16.345 13.672 16.307 / \plot 13.672 16.307 13.735 16.271 / \plot 13.735 16.271 13.794 16.233 / \plot 13.794 16.233 13.854 16.197 / \plot 13.854 16.197 13.911 16.161 / \plot 13.911 16.161 13.968 16.125 / \plot 13.968 16.125 14.021 16.091 / \plot 14.021 16.091 14.072 16.057 / \plot 14.072 16.057 14.122 16.023 / \plot 14.122 16.023 14.169 15.991 / \plot 14.169 15.991 14.216 15.960 / \plot 14.216 15.960 14.262 15.928 / \plot 14.262 15.928 14.302 15.898 / \plot 14.302 15.898 14.343 15.871 / \plot 14.343 15.871 14.381 15.841 / \plot 14.381 15.841 14.419 15.811 / \plot 14.419 15.811 14.457 15.780 / \plot 14.457 15.780 14.493 15.748 / \plot 14.493 15.748 14.531 15.714 / \plot 14.531 15.714 14.567 15.678 / \plot 14.567 15.678 14.601 15.642 / \plot 14.601 15.642 14.637 15.602 / \plot 14.637 15.602 14.671 15.562 / \plot 14.671 15.562 14.702 15.517 / \plot 14.702 15.517 14.734 15.473 / \plot 14.734 15.473 14.766 15.424 / \plot 14.766 15.424 14.796 15.371 / \plot 14.796 15.371 14.823 15.318 / \plot 14.823 15.318 14.851 15.261 / \plot 14.851 15.261 14.878 15.202 / \plot 14.878 15.202 14.901 15.138 / \plot 14.901 15.138 14.925 15.075 / \plot 14.925 15.075 14.946 15.007 / \plot 14.946 15.007 14.967 14.935 / \plot 14.967 14.935 14.986 14.863 / \plot 14.986 14.863 15.003 14.787 / \plot 15.003 14.787 15.020 14.707 / \plot 15.020 14.707 15.035 14.626 / \plot 15.035 14.626 15.047 14.539 / \plot 15.047 14.539 15.060 14.450 / \plot 15.060 14.450 15.071 14.357 / \plot 15.071 14.357 15.081 14.260 / \plot 15.081 14.260 15.088 14.186 / \plot 15.088 14.186 15.094 14.105 / \plot 15.094 14.105 15.100 14.025 / \plot 15.100 14.025 15.107 13.940 / \plot 15.107 13.940 15.111 13.854 / \plot 15.111 13.854 15.115 13.763 / \plot 15.115 13.763 15.119 13.672 / \plot 15.119 13.672 15.124 13.576 / \plot 15.124 13.576 15.128 13.477 / \plot 15.128 13.477 15.130 13.377 / \plot 15.130 13.377 15.132 13.276 / \plot 15.132 13.276 15.134 13.172 / \plot 15.134 13.172 15.136 13.066 / \plot 15.136 13.066 15.138 12.960 / \putrule from 15.138 12.960 to 15.138 12.852 \plot 15.138 12.852 15.141 12.744 / \putrule from 15.141 12.744 to 15.141 12.636 \putrule from 15.141 12.636 to 15.141 12.529 \plot 15.141 12.529 15.138 12.423 / \putrule from 15.138 12.423 to 15.138 12.315 \putrule from 15.138 12.315 to 15.138 12.211 \plot 15.138 12.211 15.136 12.107 / \plot 15.136 12.107 15.134 12.006 / \plot 15.134 12.006 15.132 11.906 / \plot 15.132 11.906 15.130 11.811 / \plot 15.130 11.811 15.128 11.716 / \plot 15.128 11.716 15.126 11.627 / \plot 15.126 11.627 15.121 11.540 / \plot 15.121 11.540 15.119 11.458 / \plot 15.119 11.458 15.117 11.379 / \plot 15.117 11.379 15.113 11.305 / \plot 15.113 11.305 15.109 11.235 / \plot 15.109 11.235 15.107 11.172 / \plot 15.107 11.172 15.102 11.110 / \plot 15.102 11.110 15.100 11.055 / \plot 15.100 11.055 15.096 11.005 / \plot 15.096 11.005 15.092 10.958 / \plot 15.092 10.958 15.088 10.916 / \plot 15.088 10.916 15.085 10.880 / \plot 15.085 10.880 15.081 10.848 / \plot 15.081 10.848 15.075 10.810 / \plot 15.075 10.810 15.071 10.780 / \plot 15.071 10.780 15.064 10.763 / \plot 15.064 10.763 15.058 10.759 / \plot 15.058 10.759 15.052 10.763 / \plot 15.052 10.763 15.045 10.778 / \plot 15.045 10.778 15.039 10.801 / \plot 15.039 10.801 15.033 10.837 / \plot 15.033 10.837 15.026 10.880 / \plot 15.026 10.880 15.018 10.933 / \plot 15.018 10.933 15.011 10.994 / \plot 15.011 10.994 15.003 11.064 / \plot 15.003 11.064 14.994 11.140 / \plot 14.994 11.140 14.986 11.223 / \plot 14.986 11.223 14.978 11.311 / \plot 14.978 11.311 14.969 11.405 / \plot 14.969 11.405 14.961 11.502 / \plot 14.961 11.502 14.950 11.599 / \plot 14.950 11.599 14.942 11.701 / \plot 14.942 11.701 14.931 11.803 / \plot 14.931 11.803 14.922 11.906 / \plot 14.922 11.906 14.912 12.008 / \plot 14.912 12.008 14.901 12.109 / \plot 14.901 12.109 14.891 12.211 / \plot 14.891 12.211 14.880 12.311 / \plot 14.880 12.311 14.870 12.408 / \plot 14.870 12.408 14.861 12.488 / \plot 14.861 12.488 14.851 12.567 / \plot 14.851 12.567 14.840 12.645 / \plot 14.840 12.645 14.829 12.725 / \plot 14.829 12.725 14.819 12.804 / \plot 14.819 12.804 14.808 12.884 / \plot 14.808 12.884 14.796 12.962 / \plot 14.796 12.962 14.783 13.043 / \plot 14.783 13.043 14.770 13.123 / \plot 14.770 13.123 14.757 13.204 / \plot 14.757 13.204 14.743 13.284 / \plot 14.743 13.284 14.728 13.367 / \plot 14.728 13.367 14.713 13.447 / \plot 14.713 13.447 14.696 13.526 / \plot 14.696 13.526 14.679 13.606 / \plot 14.679 13.606 14.662 13.684 / \plot 14.662 13.684 14.645 13.763 / \plot 14.645 13.763 14.626 13.839 / \plot 14.626 13.839 14.607 13.913 / \plot 14.607 13.913 14.590 13.985 / \plot 14.590 13.985 14.569 14.057 / \plot 14.569 14.057 14.550 14.127 / \plot 14.550 14.127 14.531 14.194 / \plot 14.531 14.194 14.510 14.260 / \plot 14.510 14.260 14.491 14.321 / \plot 14.491 14.321 14.470 14.383 / \plot 14.470 14.383 14.448 14.442 / \plot 14.448 14.442 14.427 14.499 / \plot 14.427 14.499 14.406 14.554 / \plot 14.406 14.554 14.385 14.609 / \plot 14.385 14.609 14.364 14.660 / \plot 14.364 14.660 14.340 14.711 / \plot 14.340 14.711 14.313 14.772 / \plot 14.313 14.772 14.283 14.831 / \plot 14.283 14.831 14.252 14.891 / \plot 14.252 14.891 14.220 14.948 / \plot 14.220 14.948 14.186 15.005 / \plot 14.186 15.005 14.152 15.062 / \plot 14.152 15.062 14.116 15.117 / \plot 14.116 15.117 14.080 15.174 / \plot 14.080 15.174 14.040 15.229 / \plot 14.040 15.229 14.002 15.282 / \plot 14.002 15.282 13.959 15.337 / \plot 13.959 15.337 13.919 15.390 / \plot 13.919 15.390 13.877 15.441 / \plot 13.877 15.441 13.832 15.492 / \plot 13.832 15.492 13.790 15.541 / \plot 13.790 15.541 13.748 15.589 / \plot 13.748 15.589 13.703 15.634 / \plot 13.703 15.634 13.661 15.678 / \plot 13.661 15.678 13.619 15.723 / \plot 13.619 15.723 13.576 15.763 / \plot 13.576 15.763 13.534 15.803 / \plot 13.534 15.803 13.494 15.841 / \plot 13.494 15.841 13.454 15.877 / \plot 13.454 15.877 13.413 15.913 / \plot 13.413 15.913 13.375 15.947 / \plot 13.375 15.947 13.335 15.981 / \plot 13.335 15.981 13.293 16.017 / \plot 13.293 16.017 13.250 16.053 / \plot 13.250 16.053 13.206 16.089 / \plot 13.206 16.089 13.161 16.125 / \plot 13.161 16.125 13.117 16.161 / \plot 13.117 16.161 13.070 16.197 / \plot 13.070 16.197 13.022 16.233 / \plot 13.022 16.233 12.975 16.269 / \plot 12.975 16.269 12.924 16.305 / \plot 12.924 16.305 12.874 16.339 / \plot 12.874 16.339 12.823 16.375 / \plot 12.823 16.375 12.772 16.408 / \plot 12.772 16.408 12.721 16.442 / \plot 12.721 16.442 12.668 16.474 / \plot 12.668 16.474 12.617 16.506 / \plot 12.617 16.506 12.565 16.535 / \plot 12.565 16.535 12.514 16.563 / \plot 12.514 16.563 12.463 16.590 / \plot 12.463 16.590 12.414 16.616 / \plot 12.414 16.616 12.366 16.639 / \plot 12.366 16.639 12.315 16.662 / \plot 12.315 16.662 12.268 16.684 / \plot 12.268 16.684 12.220 16.703 / \plot 12.220 16.703 12.171 16.722 / \plot 12.171 16.722 12.122 16.739 / \plot 12.122 16.739 12.071 16.756 / \plot 12.071 16.756 12.021 16.772 / \plot 12.021 16.772 11.970 16.787 / \plot 11.970 16.787 11.915 16.802 / \plot 11.915 16.802 11.860 16.815 / \plot 11.860 16.815 11.805 16.828 / \plot 11.805 16.828 11.745 16.840 / \plot 11.745 16.840 11.686 16.851 / \plot 11.686 16.851 11.625 16.861 / \plot 11.625 16.861 11.563 16.870 / \plot 11.563 16.870 11.502 16.878 / \plot 11.502 16.878 11.441 16.885 / \plot 11.441 16.885 11.377 16.889 / \plot 11.377 16.889 11.316 16.893 / \plot 11.316 16.893 11.254 16.897 / \putrule from 11.254 16.897 to 11.193 16.897 \plot 11.193 16.897 11.134 16.899 / \putrule from 11.134 16.899 to 11.074 16.899 \plot 11.074 16.899 11.017 16.897 / \plot 11.017 16.897 10.960 16.895 / \plot 10.960 16.895 10.905 16.891 / \plot 10.905 16.891 10.850 16.887 / \plot 10.850 16.887 10.795 16.880 / \plot 10.795 16.880 10.744 16.874 / \plot 10.744 16.874 10.693 16.868 / \plot 10.693 16.868 10.643 16.859 / \plot 10.643 16.859 10.590 16.851 / \plot 10.590 16.851 10.537 16.840 / \plot 10.537 16.840 10.484 16.830 / \plot 10.484 16.830 10.429 16.819 / \plot 10.429 16.819 10.372 16.806 / \plot 10.372 16.806 10.315 16.794 / \plot 10.315 16.794 10.257 16.779 / \plot 10.257 16.779 10.198 16.762 / \plot 10.198 16.762 10.139 16.747 / \plot 10.139 16.747 10.080 16.730 / \plot 10.080 16.730 10.022 16.713 / \plot 10.022 16.713 9.963 16.694 / \plot 9.963 16.694 9.904 16.675 / \plot 9.904 16.675 9.847 16.658 / \plot 9.847 16.658 9.790 16.639 / \plot 9.790 16.639 9.732 16.618 / \plot 9.732 16.618 9.677 16.599 / \plot 9.677 16.599 9.624 16.580 / \plot 9.624 16.580 9.572 16.561 / \plot 9.572 16.561 9.519 16.542 / \plot 9.519 16.542 9.468 16.523 / \plot 9.468 16.523 9.417 16.504 / \plot 9.417 16.504 9.366 16.482 / \plot 9.366 16.482 9.315 16.463 / \plot 9.315 16.463 9.265 16.444 / \plot 9.265 16.444 9.214 16.423 / \plot 9.214 16.423 9.161 16.402 / \plot 9.161 16.402 9.108 16.381 / \plot 9.108 16.381 9.053 16.358 / \plot 9.053 16.358 8.998 16.334 / \plot 8.998 16.334 8.943 16.309 / \plot 8.943 16.309 8.886 16.284 / \plot 8.886 16.284 8.827 16.258 / \plot 8.827 16.258 8.767 16.231 / \plot 8.767 16.231 8.708 16.203 / \plot 8.708 16.203 8.649 16.173 / \plot 8.649 16.173 8.587 16.146 / \plot 8.587 16.146 8.528 16.116 / \plot 8.528 16.116 8.469 16.085 / \plot 8.469 16.085 8.410 16.055 / \plot 8.410 16.055 8.350 16.023 / \plot 8.350 16.023 8.293 15.994 / \plot 8.293 15.994 8.236 15.962 / \plot 8.236 15.962 8.181 15.930 / \plot 8.181 15.930 8.126 15.898 / \plot 8.126 15.898 8.071 15.867 / \plot 8.071 15.867 8.018 15.835 / \plot 8.018 15.835 7.965 15.803 / \plot 7.965 15.803 7.912 15.769 / \plot 7.912 15.769 7.861 15.737 / \plot 7.861 15.737 7.813 15.706 / \plot 7.813 15.706 7.762 15.672 / \plot 7.762 15.672 7.709 15.638 / \plot 7.709 15.638 7.658 15.602 / \plot 7.658 15.602 7.605 15.564 / \plot 7.605 15.564 7.550 15.526 / \plot 7.550 15.526 7.495 15.486 / \plot 7.495 15.486 7.440 15.445 / \plot 7.440 15.445 7.383 15.405 / \plot 7.383 15.405 7.326 15.361 / \plot 7.326 15.361 7.269 15.318 / \plot 7.269 15.318 7.211 15.274 / \plot 7.211 15.274 7.154 15.229 / \plot 7.154 15.229 7.097 15.185 / \plot 7.097 15.185 7.042 15.138 / \plot 7.042 15.138 6.985 15.094 / \plot 6.985 15.094 6.932 15.050 / \plot 6.932 15.050 6.877 15.005 / \plot 6.877 15.005 6.826 14.961 / \plot 6.826 14.961 6.775 14.916 / \plot 6.775 14.916 6.725 14.874 / \plot 6.725 14.874 6.676 14.831 / \plot 6.676 14.831 6.629 14.791 / \plot 6.629 14.791 6.585 14.751 / \plot 6.585 14.751 6.541 14.711 / \plot 6.541 14.711 6.498 14.671 / \plot 6.498 14.671 6.456 14.630 / \plot 6.456 14.630 6.411 14.590 / \plot 6.411 14.590 6.367 14.548 / \plot 6.367 14.548 6.325 14.506 / \plot 6.325 14.506 6.280 14.463 / \plot 6.280 14.463 6.238 14.419 / \plot 6.238 14.419 6.195 14.376 / \plot 6.195 14.376 6.151 14.332 / \plot 6.151 14.332 6.109 14.287 / \plot 6.109 14.287 6.066 14.243 / \plot 6.066 14.243 6.026 14.196 / \plot 6.026 14.196 5.984 14.152 / \plot 5.984 14.152 5.944 14.105 / \plot 5.944 14.105 5.903 14.061 / \plot 5.903 14.061 5.863 14.014 / \plot 5.863 14.014 5.827 13.972 / \plot 5.827 13.972 5.789 13.928 / \plot 5.789 13.928 5.755 13.885 / \plot 5.755 13.885 5.721 13.843 / \plot 5.721 13.843 5.690 13.803 / \plot 5.690 13.803 5.658 13.763 / \plot 5.658 13.763 5.628 13.724 / \plot 5.628 13.724 5.601 13.686 / \plot 5.601 13.686 5.575 13.650 / \plot 5.575 13.650 5.550 13.614 / \plot 5.550 13.614 5.527 13.581 / \plot 5.527 13.581 5.503 13.547 / \plot 5.503 13.547 5.491 13.528 / \plot 5.491 13.528 5.478 13.509 / \plot 5.478 13.509 5.467 13.490 / \plot 5.467 13.490 5.455 13.473 / \plot 5.455 13.473 5.442 13.454 / \plot 5.442 13.454 5.431 13.437 / \plot 5.431 13.437 5.421 13.420 / \plot 5.421 13.420 5.408 13.403 / \plot 5.408 13.403 5.397 13.388 / \plot 5.397 13.388 5.387 13.373 / \plot 5.387 13.373 5.376 13.358 / \plot 5.376 13.358 5.366 13.346 / \plot 5.366 13.346 5.355 13.333 / \plot 5.355 13.333 5.345 13.322 / \plot 5.345 13.322 5.334 13.314 / \plot 5.334 13.314 5.326 13.305 / \plot 5.326 13.305 5.315 13.299 / \plot 5.315 13.299 5.306 13.293 / \plot 5.306 13.293 5.296 13.291 / \plot 5.296 13.291 5.287 13.288 / \putrule from 5.287 13.288 to 5.279 13.288 \plot 5.279 13.288 5.268 13.291 / \plot 5.268 13.291 5.260 13.295 / \plot 5.260 13.295 5.251 13.301 / \plot 5.251 13.301 5.245 13.310 / \plot 5.245 13.310 5.237 13.320 / \plot 5.237 13.320 5.228 13.335 / \plot 5.228 13.335 5.222 13.350 / \plot 5.222 13.350 5.215 13.367 / \plot 5.215 13.367 5.209 13.388 / \plot 5.209 13.388 5.203 13.411 / \plot 5.203 13.411 5.196 13.437 / \plot 5.196 13.437 5.190 13.464 / \plot 5.190 13.464 5.184 13.496 / \plot 5.184 13.496 5.177 13.528 / \plot 5.177 13.528 5.173 13.564 / \plot 5.173 13.564 5.169 13.604 / \plot 5.169 13.604 5.163 13.644 / \plot 5.163 13.644 5.158 13.688 / \plot 5.158 13.688 5.154 13.735 / \plot 5.154 13.735 5.150 13.784 / \plot 5.150 13.784 5.148 13.837 / \plot 5.148 13.837 5.143 13.892 / \plot 5.143 13.892 5.139 13.949 / \plot 5.139 13.949 5.137 14.010 / \plot 5.137 14.010 5.133 14.076 / \plot 5.133 14.076 5.131 14.144 / \plot 5.131 14.144 5.127 14.216 / \plot 5.127 14.216 5.124 14.292 / \plot 5.124 14.292 5.120 14.370 / \plot 5.120 14.370 5.118 14.453 / \plot 5.118 14.453 5.116 14.537 / \plot 5.116 14.537 5.114 14.628 / \plot 5.114 14.628 5.112 14.721 / \plot 5.112 14.721 5.110 14.817 / \plot 5.110 14.817 5.108 14.918 / \plot 5.108 14.918 5.105 15.022 / \plot 5.105 15.022 5.103 15.128 / \plot 5.103 15.128 5.101 15.238 / \plot 5.101 15.238 5.099 15.350 / \plot 5.099 15.350 5.097 15.464 / \putrule from 5.097 15.464 to 5.097 15.583 \plot 5.097 15.583 5.095 15.701 / \plot 5.095 15.701 5.093 15.824 / \putrule from 5.093 15.824 to 5.093 15.947 \plot 5.093 15.947 5.091 16.072 / \plot 5.091 16.072 5.088 16.199 / \putrule from 5.088 16.199 to 5.088 16.326 \plot 5.088 16.326 5.086 16.453 / \putrule from 5.086 16.453 to 5.086 16.580 \putrule from 5.086 16.580 to 5.086 16.709 \plot 5.086 16.709 5.084 16.836 / \putrule from 5.084 16.836 to 5.084 16.961 \putrule from 5.084 16.961 to 5.084 17.088 \plot 5.084 17.088 5.082 17.213 / \putrule from 5.082 17.213 to 5.082 17.335 \putrule from 5.082 17.335 to 5.082 17.456 \putrule from 5.082 17.456 to 5.082 17.575 \putrule from 5.082 17.575 to 5.082 17.691 \plot 5.082 17.691 5.080 17.805 / \putrule from 5.080 17.805 to 5.080 17.918 \putrule from 5.080 17.918 to 5.080 18.026 \putrule from 5.080 18.026 to 5.080 18.131 \putrule from 5.080 18.131 to 5.080 18.235 \putrule from 5.080 18.235 to 5.080 18.335 \putrule from 5.080 18.335 to 5.080 18.432 \putrule from 5.080 18.432 to 5.080 18.525 \putrule from 5.080 18.525 to 5.080 18.616 \putrule from 5.080 18.616 to 5.080 18.703 \putrule from 5.080 18.703 to 5.080 18.785 \putrule from 5.080 18.785 to 5.080 18.866 \putrule from 5.080 18.866 to 5.080 18.944 \putrule from 5.080 18.944 to 5.080 19.003 \putrule from 5.080 19.003 to 5.080 19.061 \putrule from 5.080 19.061 to 5.080 19.116 \putrule from 5.080 19.116 to 5.080 19.171 \putrule from 5.080 19.171 to 5.080 19.224 \putrule from 5.080 19.224 to 5.080 19.272 \putrule from 5.080 19.272 to 5.080 19.323 \putrule from 5.080 19.323 to 5.080 19.370 \putrule from 5.080 19.370 to 5.080 19.416 \putrule from 5.080 19.416 to 5.080 19.461 \putrule from 5.080 19.461 to 5.080 19.503 \putrule from 5.080 19.503 to 5.080 19.545 \putrule from 5.080 19.545 to 5.080 19.583 \putrule from 5.080 19.583 to 5.080 19.622 \putrule from 5.080 19.622 to 5.080 19.660 \plot 5.080 19.660 5.082 19.693 / \putrule from 5.082 19.693 to 5.082 19.727 \putrule from 5.082 19.727 to 5.082 19.761 \putrule from 5.082 19.761 to 5.082 19.791 \putrule from 5.082 19.791 to 5.082 19.820 \putrule from 5.082 19.820 to 5.082 19.848 \plot 5.082 19.848 5.084 19.873 / \putrule from 5.084 19.873 to 5.084 19.897 \putrule from 5.084 19.897 to 5.084 19.920 \putrule from 5.084 19.920 to 5.084 19.941 \putrule from 5.084 19.941 to 5.084 19.960 \plot 5.084 19.960 5.086 19.979 / \putrule from 5.086 19.979 to 5.086 19.994 \putrule from 5.086 19.994 to 5.086 20.009 \plot 5.086 20.009 5.088 20.024 / \putrule from 5.088 20.024 to 5.088 20.034 \plot 5.088 20.034 5.091 20.045 / \putrule from 5.091 20.045 to 5.091 20.053 \plot 5.091 20.053 5.093 20.060 / \putrule from 5.093 20.060 to 5.093 20.066 \plot 5.093 20.066 5.095 20.070 / \putrule from 5.095 20.070 to 5.095 20.072 \plot 5.095 20.072 5.097 20.074 / \plot 5.097 20.074 5.099 20.072 / \plot 5.099 20.072 5.101 20.070 / \putrule from 5.101 20.070 to 5.101 20.066 \plot 5.101 20.066 5.103 20.062 / \plot 5.103 20.062 5.105 20.055 / \putrule from 5.105 20.055 to 5.105 20.049 \plot 5.105 20.049 5.108 20.041 / \plot 5.108 20.041 5.110 20.030 / \plot 5.110 20.030 5.112 20.019 / \plot 5.112 20.019 5.114 20.009 / \plot 5.114 20.009 5.116 19.996 / \plot 5.116 19.996 5.118 19.983 / \plot 5.118 19.983 5.120 19.971 / \plot 5.120 19.971 5.122 19.956 / \plot 5.122 19.956 5.124 19.939 / \plot 5.124 19.939 5.127 19.924 / \plot 5.127 19.924 5.129 19.907 / \plot 5.129 19.907 5.131 19.888 / \plot 5.131 19.888 5.133 19.871 / \putrule from 5.133 19.871 to 5.133 19.869 \plot 5.133 19.869 5.137 19.833 / \plot 5.137 19.833 5.143 19.793 / \plot 5.143 19.793 5.150 19.748 / \plot 5.150 19.748 5.154 19.702 / \plot 5.154 19.702 5.163 19.653 / \plot 5.163 19.653 5.169 19.600 / \plot 5.169 19.600 5.175 19.545 / \plot 5.175 19.545 5.184 19.486 / \plot 5.184 19.486 5.192 19.425 / \plot 5.192 19.425 5.199 19.361 / \plot 5.199 19.361 5.209 19.296 / \plot 5.209 19.296 5.218 19.226 / \plot 5.218 19.226 5.226 19.156 / \plot 5.226 19.156 5.237 19.084 / \plot 5.237 19.084 5.245 19.010 / \plot 5.245 19.010 5.256 18.936 / \plot 5.256 18.936 5.266 18.860 / \plot 5.266 18.860 5.277 18.783 / \plot 5.277 18.783 5.287 18.709 / \plot 5.287 18.709 5.298 18.633 / \plot 5.298 18.633 5.306 18.559 / \plot 5.306 18.559 5.317 18.485 / \plot 5.317 18.485 5.328 18.411 / \plot 5.328 18.411 5.338 18.339 / \plot 5.338 18.339 5.349 18.269 / \plot 5.349 18.269 5.359 18.201 / \plot 5.359 18.201 5.368 18.133 / \plot 5.368 18.133 5.378 18.068 / \plot 5.378 18.068 5.387 18.002 / \plot 5.387 18.002 5.397 17.939 / }% \put{\SetFigFont{6}{7.2}{\rmdefault}{\mddefault}{\updefault}{\color[rgb]{0,0,0}$\tilde{S}_-$}% } [lB] at 4.0 12.700 \put{\SetFigFont{6}{7.2}{\rmdefault}{\mddefault}{\itdefault}{\color[rgb]{0,0,0}$\tilde{S}_+$}% } [lB] at 15.558 15.240 \put{\SetFigFont{6}{7.2}{\rmdefault}{\mddefault}{\itdefault}{\color[rgb]{0,0,0}$\bar{N}_-$}% } [lB] at 4.0 20.320 \put{\SetFigFont{6}{7.2}{\rmdefault}{\mddefault}{\itdefault}{\color[rgb]{0,0,0}$\bar{N}_+$}% } [lB] at 15.558 17.939 \put{\SetFigFont{6}{7.2}{\rmdefault}{\mddefault}{\itdefault}{\color[rgb]{0,0,0}$\tilde{N}_+$}% } [lB] at 15.558 8.096 \put{\SetFigFont{6}{7.2}{\rmdefault}{\mddefault}{\itdefault}{\color[rgb]{0,0,0}$\tilde{N}_-$}% } [lB] at 4.0 9.842 \put{\SetFigFont{6}{7.2}{\rmdefault}{\mddefault}{\itdefault}{\color[rgb]{0,0,0}$0$}% } [lB] at 4.604 16.351 \put{\SetFigFont{6}{7.2}{\rmdefault}{\mddefault}{\itdefault}{\color[rgb]{0,0,0}$-\pi$}% } [lB] at 4.0 11.271 \put{\SetFigFont{6}{7.2}{\rmdefault}{\mddefault}{\itdefault}{\color[rgb]{0,0,0}$\pi$}% } [lB] at 4.445 21.431 \put{\SetFigFont{6}{7.2}{\rmdefault}{\mddefault}{\itdefault}{\color[rgb]{0,0,0}$\mathcal{W}^-$}% } [lB] at 7.144 14.922 \put{\SetFigFont{6}{7.2}{\rmdefault}{\mddefault}{\itdefault}{\color[rgb]{0,0,0}$\mathcal{W}^+$}% } [lB] at 12.383 17.304 \put{\SetFigFont{6}{7.2}{\rmdefault}{\mddefault}{\itdefault}{\color[rgb]{0,0,0}${\mathcal K}_1$}% } [lB] at 5.715 15.399 \linethickness=0pt \putrectangle corners at 4.413 21.768 and 15.589 6.280 \endpicture} \end{center} \caption{\label{fig:CORRIDORone} The corridor ${\mathcal K}_1$ in $\mathcal{C}$ and in $\bar{\mathcal{C}}$.} \end{figure} Furthermore, in the universal cover $\bar{\mathcal{C}}$, using the well-orderedness of ${\mathbb R}$ it is possible to speak of $\widetilde{\mathcal{W}}^+$ as being situated ``above'' or ``below" $\widetilde{\mathcal{W}}^-$. It is evident that the boundary of the corridor ${\mathcal K}_1$ is oriented clockwise if $\widetilde{\mathcal{W}}^+$ is below $\widetilde{\mathcal{W}}^-$, and counterclockwise if it is the other way around. see Figure~\ref{fig:CORRIDORone}. \subsubsection{Parameter-dependent flows} Suppose that the function $g$ in \refeq{eq:flow} depends smoothly on a parameter $\mu \in I$ where $I$ is an open interval in ${\mathbb R}$ (It's enough for the $\mu$-dependence to be $C^1$). Thus the flow is now \begin{equation}\label{eq:flowmu} \left\{\begin{array}{rcl} \dot{x} & = & f(x)\\ \dot{y} & = & g_\mu(x,y).\end{array}\right. \end{equation} By the implicit function theorem, the locations of the equilibria $S_\pm,N_\pm$ also depend --in a $C^1$ fashion-- on $\mu$, so long as the non-degeneracy conditions \refeq{cond:hyp} are satisfied. We need to make certain assumptions about the $\mu$-dependence of the flow regarding its monotonicity, and the topology of its nullclines: \subsubsubsection{Monotonicity} We will only consider parameter-dependent flows for which the function $g_\mu(x,y)$ is {\em monotone non-increasing} in $\mu$. {\bf Assumption (M)}: For all $(x,y) \in \bar{\mathcal{C}}$, we have \begin{equation}\label{ass:mon} \frac{\partial}{\partial\mu} g_{\mu}(x,y) \leq 0. \end{equation} This assumption in particular implies a corresponding monotonicity for the distinguished orbits of the flow \refeq{eq:flowmu}: \begin{lem}\label{lem:mon} Let $\tilde{\mathcal{W}}^\pm_\mu = \{(x(\tau),y^\pm_\mu(\tau)\}_{\tau\in{\mathbb R}}$ be the distinguished orbits of \refeq{eq:flowmu}. Then $y^\pm_\mu$ are monotone in $\mu$, i.e. \begin{equation}\label{orbmon} \mu_1<\mu_2 \implies y^-_{\mu_1}(\tau) \geq y^-_{\mu_2}(\tau),\quad y^+_{\mu_1}(\tau) \leq y^+_{\mu_2}(\tau)\mbox{for all }\tau\in{\mathbb R}. \end{equation} \end{lem} \begin{proof} Let $$ z(\tau) := \frac{\partial}{\partial \mu} y^-_\mu(\tau).$$ Then $z$ satisfies the ODE \begin{equation}\label{odez} \dot{z} = \frac{\partial g_\mu}{\partial y}(x,y) z + \frac{\partial g_\mu}{\partial \mu}(x,y). \end{equation} By the Implicit Function Theorem and the hypotheses above, $$ z(-\infty) = \frac{ds^-(\mu)}{d\mu} = - \left.\frac{ \partial g/\partial y}{\partial g /\partial \mu}\right|_{x=x_-,y = s_-} < 0 $$ and \begin{equation}\label{odiz} \dot{z} - \frac{\partial g_\mu}{\partial y}(x,y) z \leq 0. \end{equation} Let $$ U(\tau_0,\tau) := \exp\left(-\int_{\tau_0}^\tau \frac{\partial g_\mu}{\partial y}(x(\tau'),y(\tau')) d\tau' \right) \geq 0. $$ Integrating \refeq{odiz} on $[\tau_0,\tau]$ we then obtain $$ U(\tau_0,\tau) z(\tau) \leq z(\tau_0), $$ and taking the limit $\tau_0 \to -\infty$ we conclude $z(\tau)\leq 0$ for all $\tau \in {\mathbb R}$. Therefore $y^-_\mu$ is monotone non-increasing in $\mu$. The proof of monotonicity of $y^+_\mu$ is completely analogous. \end{proof} \subsubsubsection{Topology of nullclines} Consider the subset of the cylinder $\mathcal{C}$ defined by \begin{equation} \Gamma := \{(x,y)\in \mathcal{C}\ | \ g_\mu(x,y) = 0\}. \end{equation} Thus $\Gamma$ is the zero level-set of $g_\mu$. Since $g_\mu$ is smooth, $\Gamma$ is a curve (or collection of curves) in $\mathcal{C}$. These curves are referred to as the {\em $y$-nullclines} of the flow \refeq{eq:flow}. They have the property that any orbit of the flow that crosses them must have a horizontal tangent at the crossing point. Moreover, $\Gamma$ divides $\mathcal{C}$ into two regions, thus $\mathcal{C} = \Gamma\cup{\mathcal N}\cup{\mathcal P}$, where \begin{equation} {\mathcal N} := \{(x,y)\in\mathcal{C}\ |\ g_\mu(x,y) < 0\},\qquad {\mathcal P} := \{(x,y)\ |\ g_\mu(x,y)>0\}. \end{equation} Thus the $y$ coordinate of (the lift to the universal cover of) any orbit must decrease in ${\mathcal N}$ and must increase in ${\mathcal P}$. Evidently, $\Gamma$ must include all the singular points of the flow: $S_\pm\in\Gamma$, $N_\pm \in \Gamma$. It may happen that as $\mu$ crosses a critical value $\mu^{}_c\in I$, the topology of the nullclines undergoes a dramatic change. This is indeed the case for the flows that we are studying in this paper. We introduce an assumption that amounts to having some control on this change in nullcline topology: {\bf Assumption (A)}: There exists a $\mu^{}_c \in I$ such that $\Gamma$, the zero level-set of $g_{\mu^{}_c}$ has a {\em saddle point} at some interior point of $\mathcal{C}$. In particular, for $\mu<\mu^{}_c$, the sets ${\mathcal N}$ and ${\mathcal P}$ are both connected, and $\Gamma$ is the union of two disjoint curves $\Gamma = \Gamma_u \cup \Gamma_d$, with $N_-,S_+ \in \overline{\Gamma_d}$ and $S_-,N_+\in \overline{\Gamma_u}$. On the other hand for $\mu>\mu^{}_c$, ${\mathcal N}$ is connected, while ${\mathcal P}$ has two connected components ${\mathcal P}_l$ and ${\mathcal P}_r$, each being a convex subset of $\mathcal{C}$. Moreover, $\Gamma$ is the union of two disjoint curves $\Gamma = \Gamma_l \cup \Gamma_r$, with $N_-,S_-\in\overline{\Gamma_l}$ and $ S_+,N_+\in\overline{\Gamma_r}$; see Figure~\ref{fig:topnul}. \begin{figure}[ht] \begin{center} \font\thinlinefont=cmr5 \begingroup\makeatletter\ifx\SetFigFont\undefined% \gdef\SetFigFont#1#2#3#4#5{% \reset@font\fontsize{#1}{#2pt}% \fontfamily{#3}\fontseries{#4}\fontshape{#5}% \selectfont}% \fi\endgroup% \mbox{\beginpicture \setcoordinatesystem units <0.50000cm,0.50000cm> \unitlength=0.50000cm \linethickness=1pt \setplotsymbol ({\makebox(0,0)[l]{\tencirc\symbol{'160}}}) \setshadesymbol ({\thinlinefont .}) \setlinear \linethickness=1pt \setplotsymbol ({\makebox(0,0)[l]{\tencirc\symbol{'160}}}) {\color[rgb]{0,0,0}\putrule from 5.080 21.590 to 15.240 21.590 }% \linethickness= 0.500pt \setplotsymbol ({\thinlinefont .}) {\color[rgb]{0,0,0}\putrule from 5.080 16.510 to 15.240 16.510 }% \linethickness= 0.500pt \setplotsymbol ({\thinlinefont .}) {\color[rgb]{0,0,0}\putrule from 5.080 11.430 to 15.240 11.430 }% \linethickness=1pt \setplotsymbol ({\makebox(0,0)[l]{\tencirc\symbol{'160}}}) {\color[rgb]{0,0,0}\putrule from 15.240 21.590 to 15.240 6.315 \putrule from 15.240 6.350 to 5.045 6.350 \putrule from 5.080 6.350 to 5.080 21.590 }% \linethickness= 0.500pt \setplotsymbol ({\thinlinefont .}) \setdashes < 0.1905cm> {\color[rgb]{0,0,0}\plot 15.240 15.240 15.236 15.240 / \plot 15.236 15.240 15.223 15.238 / \plot 15.223 15.238 15.202 15.234 / \plot 15.202 15.234 15.170 15.229 / \plot 15.170 15.229 15.124 15.223 / \plot 15.124 15.223 15.064 15.215 / \plot 15.064 15.215 14.994 15.204 / \plot 14.994 15.204 14.910 15.193 / \plot 14.910 15.193 14.817 15.179 / \plot 14.817 15.179 14.717 15.164 / \plot 14.717 15.164 14.611 15.149 / \plot 14.611 15.149 14.503 15.134 / \plot 14.503 15.134 14.393 15.117 / \plot 14.393 15.117 14.285 15.102 / \plot 14.285 15.102 14.180 15.085 / \plot 14.180 15.085 14.076 15.071 / \plot 14.076 15.071 13.978 15.056 / \plot 13.978 15.056 13.885 15.043 / \plot 13.885 15.043 13.796 15.030 / \plot 13.796 15.030 13.712 15.018 / \plot 13.712 15.018 13.631 15.005 / \plot 13.631 15.005 13.555 14.994 / \plot 13.555 14.994 13.483 14.982 / \plot 13.483 14.982 13.413 14.971 / \plot 13.413 14.971 13.348 14.963 / \plot 13.348 14.963 13.282 14.952 / \plot 13.282 14.952 13.219 14.942 / \plot 13.219 14.942 13.157 14.933 / \plot 13.157 14.933 13.098 14.922 / \plot 13.098 14.922 13.032 14.912 / \plot 13.032 14.912 12.967 14.901 / \plot 12.967 14.901 12.903 14.891 / \plot 12.903 14.891 12.838 14.880 / \plot 12.838 14.880 12.772 14.870 / \plot 12.772 14.870 12.706 14.857 / \plot 12.706 14.857 12.639 14.844 / \plot 12.639 14.844 12.571 14.831 / \plot 12.571 14.831 12.503 14.819 / \plot 12.503 14.819 12.435 14.806 / \plot 12.435 14.806 12.366 14.791 / \plot 12.366 14.791 12.298 14.776 / \plot 12.298 14.776 12.228 14.762 / \plot 12.228 14.762 12.158 14.745 / \plot 12.158 14.745 12.090 14.730 / \plot 12.090 14.730 12.023 14.713 / \plot 12.023 14.713 11.955 14.694 / \plot 11.955 14.694 11.889 14.677 / \plot 11.889 14.677 11.826 14.658 / \plot 11.826 14.658 11.762 14.639 / \plot 11.762 14.639 11.699 14.620 / \plot 11.699 14.620 11.637 14.601 / \plot 11.637 14.601 11.578 14.582 / \plot 11.578 14.582 11.521 14.561 / \plot 11.521 14.561 11.464 14.539 / \plot 11.464 14.539 11.407 14.518 / \plot 11.407 14.518 11.352 14.495 / \plot 11.352 14.495 11.299 14.472 / \plot 11.299 14.472 11.246 14.450 / \plot 11.246 14.450 11.195 14.427 / \plot 11.195 14.427 11.144 14.404 / \plot 11.144 14.404 11.093 14.379 / \plot 11.093 14.379 11.041 14.351 / \plot 11.041 14.351 10.988 14.323 / \plot 10.988 14.323 10.935 14.294 / \plot 10.935 14.294 10.880 14.262 / \plot 10.880 14.262 10.827 14.230 / \plot 10.827 14.230 10.772 14.196 / \plot 10.772 14.196 10.715 14.163 / \plot 10.715 14.163 10.660 14.127 / \plot 10.660 14.127 10.605 14.089 / \plot 10.605 14.089 10.547 14.048 / \plot 10.547 14.048 10.492 14.008 / \plot 10.492 14.008 10.437 13.968 / \plot 10.437 13.968 10.382 13.926 / \plot 10.382 13.926 10.327 13.883 / \plot 10.327 13.883 10.274 13.839 / \plot 10.274 13.839 10.224 13.796 / \plot 10.224 13.796 10.171 13.752 / \plot 10.171 13.752 10.122 13.705 / \plot 10.122 13.705 10.073 13.661 / \plot 10.073 13.661 10.025 13.617 / \plot 10.025 13.617 9.980 13.570 / \plot 9.980 13.570 9.934 13.523 / \plot 9.934 13.523 9.889 13.477 / \plot 9.889 13.477 9.847 13.430 / \plot 9.847 13.430 9.804 13.384 / \plot 9.804 13.384 9.764 13.335 / \plot 9.764 13.335 9.724 13.288 / \plot 9.724 13.288 9.686 13.242 / \plot 9.686 13.242 9.648 13.193 / \plot 9.648 13.193 9.610 13.144 / \plot 9.610 13.144 9.572 13.094 / \plot 9.572 13.094 9.533 13.043 / \plot 9.533 13.043 9.495 12.988 / \plot 9.495 12.988 9.455 12.933 / \plot 9.455 12.933 9.415 12.878 / \plot 9.415 12.878 9.377 12.821 / \plot 9.377 12.821 9.337 12.761 / \plot 9.337 12.761 9.296 12.702 / \plot 9.296 12.702 9.256 12.641 / \plot 9.256 12.641 9.216 12.579 / \plot 9.216 12.579 9.176 12.518 / \plot 9.176 12.518 9.136 12.454 / \plot 9.136 12.454 9.095 12.391 / \plot 9.095 12.391 9.055 12.330 / \plot 9.055 12.330 9.017 12.268 / \plot 9.017 12.268 8.977 12.205 / \plot 8.977 12.205 8.939 12.143 / \plot 8.939 12.143 8.901 12.084 / \plot 8.901 12.084 8.862 12.025 / \plot 8.862 12.025 8.827 11.966 / \plot 8.827 11.966 8.788 11.908 / \plot 8.788 11.908 8.752 11.851 / \plot 8.752 11.851 8.719 11.796 / \plot 8.719 11.796 8.683 11.741 / \plot 8.683 11.741 8.649 11.688 / \plot 8.649 11.688 8.615 11.637 / \plot 8.615 11.637 8.581 11.587 / \plot 8.581 11.587 8.547 11.536 / \plot 8.547 11.536 8.507 11.479 / \plot 8.507 11.479 8.467 11.422 / \plot 8.467 11.422 8.426 11.366 / \plot 8.426 11.366 8.386 11.309 / \plot 8.386 11.309 8.344 11.254 / \plot 8.344 11.254 8.302 11.197 / \plot 8.302 11.197 8.259 11.142 / \plot 8.259 11.142 8.215 11.085 / \plot 8.215 11.085 8.170 11.028 / \plot 8.170 11.028 8.124 10.973 / \plot 8.124 10.973 8.077 10.916 / \plot 8.077 10.916 8.029 10.861 / \plot 8.029 10.861 7.982 10.806 / \plot 7.982 10.806 7.933 10.753 / \plot 7.933 10.753 7.887 10.700 / \plot 7.887 10.700 7.838 10.649 / \plot 7.838 10.649 7.789 10.598 / \plot 7.789 10.598 7.743 10.549 / \plot 7.743 10.549 7.696 10.503 / \plot 7.696 10.503 7.650 10.458 / \plot 7.650 10.458 7.603 10.414 / \plot 7.603 10.414 7.559 10.374 / \plot 7.559 10.374 7.514 10.334 / \plot 7.514 10.334 7.472 10.295 / \plot 7.472 10.295 7.427 10.259 / \plot 7.427 10.259 7.385 10.226 / \plot 7.385 10.226 7.345 10.192 / \plot 7.345 10.192 7.303 10.160 / \plot 7.303 10.160 7.254 10.124 / \plot 7.254 10.124 7.205 10.090 / \plot 7.205 10.090 7.156 10.056 / \plot 7.156 10.056 7.106 10.025 / \plot 7.106 10.025 7.055 9.993 / \plot 7.055 9.993 7.002 9.963 / \plot 7.002 9.963 6.949 9.934 / \plot 6.949 9.934 6.894 9.906 / \plot 6.894 9.906 6.841 9.881 / \plot 6.841 9.881 6.786 9.855 / \plot 6.786 9.855 6.731 9.832 / \plot 6.731 9.832 6.674 9.811 / \plot 6.674 9.811 6.621 9.792 / \plot 6.621 9.792 6.566 9.775 / \plot 6.566 9.775 6.513 9.758 / \plot 6.513 9.758 6.460 9.745 / \plot 6.460 9.745 6.407 9.735 / \plot 6.407 9.735 6.358 9.724 / \plot 6.358 9.724 6.310 9.718 / \plot 6.310 9.718 6.263 9.711 / \plot 6.263 9.711 6.217 9.709 / \plot 6.217 9.709 6.172 9.707 / \plot 6.172 9.707 6.128 9.707 / \plot 6.128 9.707 6.085 9.709 / \plot 6.085 9.709 6.039 9.716 / \plot 6.039 9.716 5.994 9.722 / \plot 5.994 9.722 5.948 9.730 / \plot 5.948 9.730 5.901 9.743 / \plot 5.901 9.743 5.853 9.758 / \plot 5.853 9.758 5.802 9.775 / \plot 5.802 9.775 5.749 9.796 / \plot 5.749 9.796 5.692 9.821 / \plot 5.692 9.821 5.632 9.849 / \plot 5.632 9.849 5.569 9.881 / \plot 5.569 9.881 5.503 9.914 / \plot 5.503 9.914 5.438 9.950 / \plot 5.438 9.950 5.370 9.989 / \plot 5.370 9.989 5.306 10.025 / \plot 5.306 10.025 5.247 10.061 / \plot 5.247 10.061 5.194 10.090 / \plot 5.194 10.090 5.152 10.116 / \plot 5.152 10.116 5.120 10.137 / \plot 5.120 10.137 5.097 10.149 / \plot 5.097 10.149 5.086 10.156 / \plot 5.086 10.156 5.080 10.160 / }% \linethickness= 0.500pt \setplotsymbol ({\thinlinefont .}) {\color[rgb]{0,0,0}\plot 15.240 17.780 15.236 17.776 / \plot 15.236 17.776 15.225 17.767 / \plot 15.225 17.767 15.204 17.755 / \plot 15.204 17.755 15.174 17.731 / \plot 15.174 17.731 15.134 17.702 / \plot 15.134 17.702 15.085 17.666 / \plot 15.085 17.666 15.028 17.621 / \plot 15.028 17.621 14.965 17.573 / \plot 14.965 17.573 14.897 17.522 / \plot 14.897 17.522 14.829 17.469 / \plot 14.829 17.469 14.760 17.416 / \plot 14.760 17.416 14.694 17.363 / \plot 14.694 17.363 14.630 17.312 / \plot 14.630 17.312 14.569 17.264 / \plot 14.569 17.264 14.514 17.217 / \plot 14.514 17.217 14.459 17.173 / \plot 14.459 17.173 14.410 17.128 / \plot 14.410 17.128 14.364 17.086 / \plot 14.364 17.086 14.319 17.046 / \plot 14.319 17.046 14.277 17.005 / \plot 14.277 17.005 14.235 16.963 / \plot 14.235 16.963 14.194 16.923 / \plot 14.194 16.923 14.156 16.880 / \plot 14.156 16.880 14.122 16.844 / \plot 14.122 16.844 14.089 16.806 / \plot 14.089 16.806 14.053 16.768 / \plot 14.053 16.768 14.019 16.728 / \plot 14.019 16.728 13.985 16.686 / \plot 13.985 16.686 13.949 16.643 / \plot 13.949 16.643 13.911 16.601 / \plot 13.911 16.601 13.875 16.557 / \plot 13.875 16.557 13.835 16.510 / \plot 13.835 16.510 13.796 16.463 / \plot 13.796 16.463 13.756 16.417 / \plot 13.756 16.417 13.714 16.368 / \plot 13.714 16.368 13.674 16.320 / \plot 13.674 16.320 13.631 16.271 / \plot 13.631 16.271 13.587 16.224 / \plot 13.587 16.224 13.545 16.176 / \plot 13.545 16.176 13.500 16.127 / \plot 13.500 16.127 13.456 16.080 / \plot 13.456 16.080 13.409 16.034 / \plot 13.409 16.034 13.365 15.987 / \plot 13.365 15.987 13.320 15.943 / \plot 13.320 15.943 13.274 15.900 / \plot 13.274 15.900 13.227 15.858 / \plot 13.227 15.858 13.180 15.816 / \plot 13.180 15.816 13.134 15.776 / \plot 13.134 15.776 13.087 15.737 / \plot 13.087 15.737 13.041 15.699 / \plot 13.041 15.699 12.992 15.663 / \plot 12.992 15.663 12.943 15.629 / \plot 12.943 15.629 12.897 15.596 / \plot 12.897 15.596 12.846 15.564 / \plot 12.846 15.564 12.795 15.532 / \plot 12.795 15.532 12.742 15.498 / \plot 12.742 15.498 12.689 15.466 / \plot 12.689 15.466 12.632 15.435 / \plot 12.632 15.435 12.573 15.405 / \plot 12.573 15.405 12.514 15.373 / \plot 12.514 15.373 12.452 15.344 / \plot 12.452 15.344 12.389 15.314 / \plot 12.389 15.314 12.325 15.284 / \plot 12.325 15.284 12.260 15.255 / \plot 12.260 15.255 12.194 15.227 / \plot 12.194 15.227 12.126 15.202 / \plot 12.126 15.202 12.059 15.174 / \plot 12.059 15.174 11.991 15.151 / \plot 11.991 15.151 11.923 15.128 / \plot 11.923 15.128 11.855 15.105 / \plot 11.855 15.105 11.788 15.083 / \plot 11.788 15.083 11.722 15.064 / \plot 11.722 15.064 11.656 15.047 / \plot 11.656 15.047 11.591 15.030 / \plot 11.591 15.030 11.527 15.014 / \plot 11.527 15.014 11.464 15.001 / \plot 11.464 15.001 11.402 14.988 / \plot 11.402 14.988 11.343 14.975 / \plot 11.343 14.975 11.282 14.965 / \plot 11.282 14.965 11.225 14.956 / \plot 11.225 14.956 11.165 14.948 / \plot 11.165 14.948 11.104 14.942 / \plot 11.104 14.942 11.041 14.935 / \plot 11.041 14.935 10.979 14.931 / \plot 10.979 14.931 10.916 14.927 / \plot 10.916 14.927 10.852 14.925 / \plot 10.852 14.925 10.787 14.922 / \plot 10.787 14.922 10.721 14.922 / \plot 10.721 14.922 10.655 14.922 / \plot 10.655 14.922 10.588 14.925 / \plot 10.588 14.925 10.520 14.927 / \plot 10.520 14.927 10.450 14.931 / \plot 10.450 14.931 10.382 14.933 / \plot 10.382 14.933 10.312 14.939 / \plot 10.312 14.939 10.243 14.944 / \plot 10.243 14.944 10.173 14.950 / \plot 10.173 14.950 10.105 14.958 / \plot 10.105 14.958 10.035 14.965 / \plot 10.035 14.965 9.967 14.973 / \plot 9.967 14.973 9.902 14.980 / \plot 9.902 14.980 9.834 14.988 / \plot 9.834 14.988 9.771 14.997 / \plot 9.771 14.997 9.705 15.005 / \plot 9.705 15.005 9.644 15.014 / \plot 9.644 15.014 9.580 15.022 / \plot 9.580 15.022 9.521 15.030 / \plot 9.521 15.030 9.459 15.039 / \plot 9.459 15.039 9.400 15.047 / \plot 9.400 15.047 9.341 15.054 / \plot 9.341 15.054 9.279 15.062 / \plot 9.279 15.062 9.220 15.071 / \plot 9.220 15.071 9.159 15.077 / \plot 9.159 15.077 9.095 15.083 / \plot 9.095 15.083 9.032 15.090 / \plot 9.032 15.090 8.968 15.096 / \plot 8.968 15.096 8.903 15.102 / \plot 8.903 15.102 8.835 15.107 / \plot 8.835 15.107 8.767 15.113 / \plot 8.767 15.113 8.700 15.115 / \plot 8.700 15.115 8.630 15.117 / \plot 8.630 15.117 8.560 15.119 / \plot 8.560 15.119 8.488 15.121 / \plot 8.488 15.121 8.418 15.119 / \plot 8.418 15.119 8.346 15.119 / \plot 8.346 15.119 8.276 15.115 / \plot 8.276 15.115 8.206 15.111 / \plot 8.206 15.111 8.136 15.107 / \plot 8.136 15.107 8.069 15.100 / \plot 8.069 15.100 8.003 15.092 / \plot 8.003 15.092 7.938 15.081 / \plot 7.938 15.081 7.874 15.071 / \plot 7.874 15.071 7.811 15.060 / \plot 7.811 15.060 7.749 15.045 / \plot 7.749 15.045 7.688 15.030 / \plot 7.688 15.030 7.631 15.014 / \plot 7.631 15.014 7.571 14.994 / \plot 7.571 14.994 7.514 14.975 / \plot 7.514 14.975 7.457 14.954 / \plot 7.457 14.954 7.400 14.931 / \plot 7.400 14.931 7.343 14.906 / \plot 7.343 14.906 7.286 14.876 / \plot 7.286 14.876 7.228 14.846 / \plot 7.228 14.846 7.169 14.817 / \plot 7.169 14.817 7.110 14.783 / \plot 7.110 14.783 7.051 14.747 / \plot 7.051 14.747 6.991 14.709 / \plot 6.991 14.709 6.932 14.669 / \plot 6.932 14.669 6.871 14.628 / \plot 6.871 14.628 6.811 14.586 / \plot 6.811 14.586 6.752 14.541 / \plot 6.752 14.541 6.693 14.497 / \plot 6.693 14.497 6.636 14.450 / \plot 6.636 14.450 6.579 14.404 / \plot 6.579 14.404 6.521 14.357 / \plot 6.521 14.357 6.466 14.311 / \plot 6.466 14.311 6.414 14.262 / \plot 6.414 14.262 6.361 14.216 / \plot 6.361 14.216 6.312 14.171 / \plot 6.312 14.171 6.263 14.125 / \plot 6.263 14.125 6.217 14.080 / \plot 6.217 14.080 6.172 14.036 / \plot 6.172 14.036 6.128 13.991 / \plot 6.128 13.991 6.085 13.949 / \plot 6.085 13.949 6.045 13.906 / \plot 6.045 13.906 6.007 13.864 / \plot 6.007 13.864 5.965 13.820 / \plot 5.965 13.820 5.925 13.775 / \plot 5.925 13.775 5.884 13.731 / \plot 5.884 13.731 5.844 13.686 / \plot 5.844 13.686 5.804 13.640 / \plot 5.804 13.640 5.764 13.591 / \plot 5.764 13.591 5.723 13.542 / \plot 5.723 13.542 5.681 13.492 / \plot 5.681 13.492 5.639 13.437 / \plot 5.639 13.437 5.592 13.379 / \plot 5.592 13.379 5.546 13.320 / \plot 5.546 13.320 5.497 13.257 / \plot 5.497 13.257 5.448 13.193 / \plot 5.448 13.193 5.400 13.128 / \plot 5.400 13.128 5.349 13.062 / \plot 5.349 13.062 5.300 12.998 / \plot 5.300 12.998 5.256 12.937 / \plot 5.256 12.937 5.213 12.880 / \plot 5.213 12.880 5.175 12.829 / \plot 5.175 12.829 5.143 12.787 / \plot 5.143 12.787 5.118 12.753 / \plot 5.118 12.753 5.101 12.730 / \plot 5.101 12.730 5.088 12.713 / \plot 5.088 12.713 5.082 12.704 / \plot 5.082 12.704 5.080 12.700 / }% \linethickness= 0.500pt \setplotsymbol ({\thinlinefont .}) \setsolid {\color[rgb]{0,0,0}\plot 5.080 12.700 5.082 12.702 / \plot 5.082 12.702 5.091 12.708 / \plot 5.091 12.708 5.103 12.721 / \plot 5.103 12.721 5.122 12.738 / \plot 5.122 12.738 5.150 12.761 / \plot 5.150 12.761 5.186 12.791 / \plot 5.186 12.791 5.228 12.827 / \plot 5.228 12.827 5.277 12.867 / \plot 5.277 12.867 5.330 12.912 / \plot 5.330 12.912 5.387 12.958 / \plot 5.387 12.958 5.446 13.005 / \plot 5.446 13.005 5.508 13.051 / \plot 5.508 13.051 5.569 13.096 / \plot 5.569 13.096 5.630 13.140 / \plot 5.630 13.140 5.692 13.183 / \plot 5.692 13.183 5.753 13.223 / \plot 5.753 13.223 5.814 13.261 / \plot 5.814 13.261 5.876 13.297 / \plot 5.876 13.297 5.937 13.331 / \plot 5.937 13.331 6.001 13.365 / \plot 6.001 13.365 6.064 13.396 / \plot 6.064 13.396 6.132 13.428 / \plot 6.132 13.428 6.200 13.460 / \plot 6.200 13.460 6.274 13.490 / \plot 6.274 13.490 6.350 13.519 / \plot 6.350 13.519 6.401 13.540 / \plot 6.401 13.540 6.456 13.561 / \plot 6.456 13.561 6.511 13.581 / \plot 6.511 13.581 6.568 13.602 / \plot 6.568 13.602 6.627 13.623 / \plot 6.627 13.623 6.689 13.644 / \plot 6.689 13.644 6.752 13.665 / \plot 6.752 13.665 6.820 13.686 / \plot 6.820 13.686 6.888 13.710 / \plot 6.888 13.710 6.957 13.733 / \plot 6.957 13.733 7.032 13.754 / \plot 7.032 13.754 7.106 13.777 / \plot 7.106 13.777 7.184 13.803 / \plot 7.184 13.803 7.264 13.826 / \plot 7.264 13.826 7.345 13.849 / \plot 7.345 13.849 7.430 13.875 / \plot 7.430 13.875 7.514 13.898 / \plot 7.514 13.898 7.601 13.923 / \plot 7.601 13.923 7.690 13.949 / \plot 7.690 13.949 7.781 13.974 / \plot 7.781 13.974 7.872 14.000 / \plot 7.872 14.000 7.963 14.023 / \plot 7.963 14.023 8.056 14.048 / \plot 8.056 14.048 8.149 14.074 / \plot 8.149 14.074 8.242 14.099 / \plot 8.242 14.099 8.338 14.125 / \plot 8.338 14.125 8.431 14.148 / \plot 8.431 14.148 8.526 14.173 / \plot 8.526 14.173 8.619 14.196 / \plot 8.619 14.196 8.714 14.220 / \plot 8.714 14.220 8.807 14.243 / \plot 8.807 14.243 8.901 14.268 / \plot 8.901 14.268 8.992 14.290 / \plot 8.992 14.290 9.085 14.313 / \plot 9.085 14.313 9.176 14.336 / \plot 9.176 14.336 9.267 14.357 / \plot 9.267 14.357 9.358 14.381 / \plot 9.358 14.381 9.449 14.402 / \plot 9.449 14.402 9.540 14.425 / \plot 9.540 14.425 9.631 14.446 / \plot 9.631 14.446 9.718 14.467 / \plot 9.718 14.467 9.804 14.489 / \plot 9.804 14.489 9.893 14.510 / \plot 9.893 14.510 9.982 14.531 / \plot 9.982 14.531 10.071 14.552 / \plot 10.071 14.552 10.162 14.573 / \plot 10.162 14.573 10.253 14.594 / \plot 10.253 14.594 10.346 14.618 / \plot 10.346 14.618 10.439 14.639 / \plot 10.439 14.639 10.535 14.662 / \plot 10.535 14.662 10.630 14.685 / \plot 10.630 14.685 10.727 14.709 / \plot 10.727 14.709 10.825 14.734 / \plot 10.825 14.734 10.922 14.757 / \plot 10.922 14.757 11.021 14.783 / \plot 11.021 14.783 11.119 14.808 / \plot 11.119 14.808 11.218 14.834 / \plot 11.218 14.834 11.318 14.859 / \plot 11.318 14.859 11.417 14.884 / \plot 11.417 14.884 11.515 14.910 / \plot 11.515 14.910 11.614 14.937 / \plot 11.614 14.937 11.712 14.963 / \plot 11.712 14.963 11.807 14.990 / \plot 11.807 14.990 11.902 15.018 / \plot 11.902 15.018 11.997 15.045 / \plot 11.997 15.045 12.088 15.071 / \plot 12.088 15.071 12.181 15.098 / \plot 12.181 15.098 12.270 15.126 / \plot 12.270 15.126 12.357 15.153 / \plot 12.357 15.153 12.444 15.181 / \plot 12.444 15.181 12.526 15.206 / \plot 12.526 15.206 12.609 15.234 / \plot 12.609 15.234 12.687 15.261 / \plot 12.687 15.261 12.766 15.289 / \plot 12.766 15.289 12.840 15.314 / \plot 12.840 15.314 12.914 15.342 / \plot 12.914 15.342 12.984 15.369 / \plot 12.984 15.369 13.051 15.395 / \plot 13.051 15.395 13.119 15.422 / \plot 13.119 15.422 13.185 15.450 / \plot 13.185 15.450 13.246 15.477 / \plot 13.246 15.477 13.310 15.505 / \plot 13.310 15.505 13.392 15.543 / \plot 13.392 15.543 13.473 15.583 / \plot 13.473 15.583 13.551 15.625 / \plot 13.551 15.625 13.629 15.665 / \plot 13.629 15.665 13.703 15.710 / \plot 13.703 15.710 13.775 15.754 / \plot 13.775 15.754 13.847 15.799 / \plot 13.847 15.799 13.917 15.845 / \plot 13.917 15.845 13.985 15.892 / \plot 13.985 15.892 14.050 15.941 / \plot 14.050 15.941 14.116 15.989 / \plot 14.116 15.989 14.177 16.040 / \plot 14.177 16.040 14.237 16.091 / \plot 14.237 16.091 14.296 16.142 / \plot 14.296 16.142 14.351 16.192 / \plot 14.351 16.192 14.404 16.243 / \plot 14.404 16.243 14.455 16.294 / \plot 14.455 16.294 14.503 16.345 / \plot 14.503 16.345 14.548 16.396 / \plot 14.548 16.396 14.590 16.444 / \plot 14.590 16.444 14.633 16.493 / \plot 14.633 16.493 14.669 16.540 / \plot 14.669 16.540 14.704 16.586 / \plot 14.704 16.586 14.738 16.631 / \plot 14.738 16.631 14.768 16.675 / \plot 14.768 16.675 14.798 16.720 / \plot 14.798 16.720 14.825 16.760 / \plot 14.825 16.760 14.851 16.802 / \plot 14.851 16.802 14.874 16.842 / \plot 14.874 16.842 14.897 16.880 / \plot 14.897 16.880 14.925 16.933 / \plot 14.925 16.933 14.952 16.984 / \plot 14.952 16.984 14.975 17.035 / \plot 14.975 17.035 14.999 17.088 / \plot 14.999 17.088 15.022 17.141 / \plot 15.022 17.141 15.043 17.194 / \plot 15.043 17.194 15.062 17.251 / \plot 15.062 17.251 15.083 17.310 / \plot 15.083 17.310 15.102 17.374 / \plot 15.102 17.374 15.121 17.437 / \plot 15.121 17.437 15.138 17.505 / \plot 15.138 17.505 15.157 17.575 / \plot 15.157 17.575 15.174 17.642 / \plot 15.174 17.642 15.189 17.708 / \plot 15.189 17.708 15.204 17.769 / \plot 15.204 17.769 15.215 17.822 / \plot 15.215 17.822 15.225 17.867 / \plot 15.225 17.867 15.232 17.899 / \plot 15.232 17.899 15.236 17.922 / \plot 15.236 17.922 15.238 17.932 / \plot 15.238 17.932 15.240 17.939 / }% \linethickness= 0.500pt \setplotsymbol ({\thinlinefont .}) {\color[rgb]{0,0,0}\putrule from 15.240 15.240 to 15.236 15.240 \putrule from 15.236 15.240 to 15.227 15.240 \plot 15.227 15.240 15.208 15.238 / \plot 15.208 15.238 15.183 15.236 / \plot 15.183 15.236 15.145 15.234 / \plot 15.145 15.234 15.094 15.229 / \plot 15.094 15.229 15.033 15.225 / \plot 15.033 15.225 14.958 15.219 / \plot 14.958 15.219 14.874 15.212 / \plot 14.874 15.212 14.781 15.206 / \plot 14.781 15.206 14.679 15.198 / \plot 14.679 15.198 14.571 15.189 / \plot 14.571 15.189 14.461 15.181 / \plot 14.461 15.181 14.347 15.172 / \plot 14.347 15.172 14.232 15.162 / \plot 14.232 15.162 14.120 15.153 / \plot 14.120 15.153 14.008 15.143 / \plot 14.008 15.143 13.898 15.134 / \plot 13.898 15.134 13.792 15.124 / \plot 13.792 15.124 13.691 15.113 / \plot 13.691 15.113 13.593 15.105 / \plot 13.593 15.105 13.498 15.094 / \plot 13.498 15.094 13.407 15.085 / \plot 13.407 15.085 13.320 15.075 / \plot 13.320 15.075 13.236 15.064 / \plot 13.236 15.064 13.153 15.054 / \plot 13.153 15.054 13.075 15.045 / \plot 13.075 15.045 12.996 15.035 / \plot 12.996 15.035 12.920 15.022 / \plot 12.920 15.022 12.844 15.011 / \plot 12.844 15.011 12.770 15.001 / \plot 12.770 15.001 12.696 14.988 / \plot 12.696 14.988 12.622 14.975 / \plot 12.622 14.975 12.554 14.963 / \plot 12.554 14.963 12.486 14.952 / \plot 12.486 14.952 12.416 14.939 / \plot 12.416 14.939 12.349 14.927 / \plot 12.349 14.927 12.279 14.912 / \plot 12.279 14.912 12.207 14.897 / \plot 12.207 14.897 12.135 14.882 / \plot 12.135 14.882 12.061 14.867 / \plot 12.061 14.867 11.985 14.853 / \plot 11.985 14.853 11.908 14.836 / \plot 11.908 14.836 11.832 14.819 / \plot 11.832 14.819 11.754 14.800 / \plot 11.754 14.800 11.673 14.781 / \plot 11.673 14.781 11.593 14.762 / \plot 11.593 14.762 11.510 14.743 / \plot 11.510 14.743 11.428 14.721 / \plot 11.428 14.721 11.345 14.702 / \plot 11.345 14.702 11.261 14.681 / \plot 11.261 14.681 11.178 14.658 / \plot 11.178 14.658 11.093 14.637 / \plot 11.093 14.637 11.009 14.613 / \plot 11.009 14.613 10.924 14.592 / \plot 10.924 14.592 10.839 14.569 / \plot 10.839 14.569 10.757 14.546 / \plot 10.757 14.546 10.672 14.522 / \plot 10.672 14.522 10.590 14.497 / \plot 10.590 14.497 10.509 14.474 / \plot 10.509 14.474 10.427 14.450 / \plot 10.427 14.450 10.346 14.427 / \plot 10.346 14.427 10.268 14.402 / \plot 10.268 14.402 10.190 14.379 / \plot 10.190 14.379 10.111 14.355 / \plot 10.111 14.355 10.033 14.330 / \plot 10.033 14.330 9.957 14.307 / \plot 9.957 14.307 9.883 14.281 / \plot 9.883 14.281 9.807 14.258 / \plot 9.807 14.258 9.732 14.232 / \plot 9.732 14.232 9.658 14.207 / \plot 9.658 14.207 9.578 14.182 / \plot 9.578 14.182 9.500 14.154 / \plot 9.500 14.154 9.419 14.127 / \plot 9.419 14.127 9.339 14.099 / \plot 9.339 14.099 9.258 14.072 / \plot 9.258 14.072 9.176 14.042 / \plot 9.176 14.042 9.093 14.012 / \plot 9.093 14.012 9.009 13.983 / \plot 9.009 13.983 8.924 13.951 / \plot 8.924 13.951 8.837 13.919 / \plot 8.837 13.919 8.752 13.887 / \plot 8.752 13.887 8.664 13.856 / \plot 8.664 13.856 8.577 13.822 / \plot 8.577 13.822 8.490 13.788 / \plot 8.490 13.788 8.401 13.754 / \plot 8.401 13.754 8.314 13.720 / \plot 8.314 13.720 8.227 13.686 / \plot 8.227 13.686 8.141 13.653 / \plot 8.141 13.653 8.054 13.619 / \plot 8.054 13.619 7.969 13.585 / \plot 7.969 13.585 7.887 13.551 / \plot 7.887 13.551 7.804 13.517 / \plot 7.804 13.517 7.722 13.483 / \plot 7.722 13.483 7.643 13.449 / \plot 7.643 13.449 7.567 13.418 / \plot 7.567 13.418 7.491 13.386 / \plot 7.491 13.386 7.419 13.354 / \plot 7.419 13.354 7.347 13.322 / \plot 7.347 13.322 7.277 13.293 / \plot 7.277 13.293 7.211 13.263 / \plot 7.211 13.263 7.146 13.233 / \plot 7.146 13.233 7.084 13.206 / \plot 7.084 13.206 7.023 13.178 / \plot 7.023 13.178 6.966 13.151 / \plot 6.966 13.151 6.909 13.123 / \plot 6.909 13.123 6.854 13.096 / \plot 6.854 13.096 6.778 13.060 / \plot 6.778 13.060 6.706 13.026 / \plot 6.706 13.026 6.636 12.990 / \plot 6.636 12.990 6.568 12.954 / \plot 6.568 12.954 6.500 12.918 / \plot 6.500 12.918 6.435 12.882 / \plot 6.435 12.882 6.371 12.846 / \plot 6.371 12.846 6.308 12.810 / \plot 6.308 12.810 6.248 12.774 / \plot 6.248 12.774 6.189 12.738 / \plot 6.189 12.738 6.132 12.700 / \plot 6.132 12.700 6.075 12.664 / \plot 6.075 12.664 6.022 12.628 / \plot 6.022 12.628 5.971 12.592 / \plot 5.971 12.592 5.925 12.556 / \plot 5.925 12.556 5.878 12.520 / \plot 5.878 12.520 5.836 12.486 / \plot 5.836 12.486 5.795 12.452 / \plot 5.795 12.452 5.757 12.418 / \plot 5.757 12.418 5.721 12.385 / \plot 5.721 12.385 5.690 12.353 / \plot 5.690 12.353 5.658 12.321 / \plot 5.658 12.321 5.630 12.289 / \plot 5.630 12.289 5.605 12.260 / \plot 5.605 12.260 5.580 12.228 / \plot 5.580 12.228 5.556 12.196 / \plot 5.556 12.196 5.529 12.156 / \plot 5.529 12.156 5.501 12.116 / \plot 5.501 12.116 5.478 12.073 / \plot 5.478 12.073 5.455 12.031 / \plot 5.455 12.031 5.431 11.985 / \plot 5.431 11.985 5.410 11.938 / \plot 5.410 11.938 5.391 11.891 / \plot 5.391 11.891 5.372 11.841 / \plot 5.372 11.841 5.355 11.792 / \plot 5.355 11.792 5.340 11.741 / \plot 5.340 11.741 5.326 11.688 / \plot 5.326 11.688 5.313 11.637 / \plot 5.313 11.637 5.300 11.587 / \plot 5.300 11.587 5.290 11.536 / \plot 5.290 11.536 5.279 11.485 / \plot 5.279 11.485 5.271 11.436 / \plot 5.271 11.436 5.262 11.388 / \plot 5.262 11.388 5.254 11.339 / \plot 5.254 11.339 5.245 11.292 / \plot 5.245 11.292 5.239 11.244 / \plot 5.239 11.244 5.232 11.201 / \plot 5.232 11.201 5.226 11.157 / \plot 5.226 11.157 5.218 11.113 / \plot 5.218 11.113 5.211 11.066 / \plot 5.211 11.066 5.205 11.015 / \plot 5.205 11.015 5.196 10.962 / \plot 5.196 10.962 5.188 10.907 / \plot 5.188 10.907 5.179 10.846 / \plot 5.179 10.846 5.169 10.780 / \plot 5.169 10.780 5.160 10.712 / \plot 5.160 10.712 5.150 10.640 / \plot 5.150 10.640 5.139 10.564 / \plot 5.139 10.564 5.127 10.490 / \plot 5.127 10.490 5.118 10.418 / \plot 5.118 10.418 5.108 10.351 / \plot 5.108 10.351 5.099 10.291 / \plot 5.099 10.291 5.093 10.243 / \plot 5.093 10.243 5.086 10.204 / \plot 5.086 10.204 5.082 10.181 / \plot 5.082 10.181 5.080 10.166 / \putrule from 5.080 10.166 to 5.080 10.160 }% \linethickness= 0.500pt \setplotsymbol ({\thinlinefont .}) {\color[rgb]{0,0,0}\plot 5.112 11.430 6.636 9.874 / }% \linethickness= 0.500pt \setplotsymbol ({\thinlinefont .}) {\color[rgb]{0,0,0}\plot 5.112 12.732 7.398 10.382 / }% \linethickness= 0.500pt \setplotsymbol ({\thinlinefont .}) {\color[rgb]{0,0,0}\plot 5.620 13.399 8.064 11.081 / }% \linethickness= 0.500pt \setplotsymbol ({\thinlinefont .}) {\color[rgb]{0,0,0}\plot 6.223 14.097 8.604 11.811 / }% \linethickness= 0.500pt \setplotsymbol ({\thinlinefont .}) {\color[rgb]{0,0,0}\plot 6.890 14.637 9.049 12.541 / }% \linethickness= 0.500pt \setplotsymbol ({\thinlinefont .}) {\color[rgb]{0,0,0}\plot 7.715 15.081 9.652 13.272 / }% \linethickness= 0.500pt \setplotsymbol ({\thinlinefont .}) {\color[rgb]{0,0,0}\plot 12.522 15.369 12.975 14.956 / }% \linethickness= 0.500pt \setplotsymbol ({\thinlinefont .}) {\color[rgb]{0,0,0}\plot 13.193 15.903 13.942 15.153 / }% \linethickness= 0.500pt \setplotsymbol ({\thinlinefont .}) {\color[rgb]{0,0,0}\plot 13.843 16.612 15.145 15.291 / }% \linethickness= 0.500pt \setplotsymbol ({\thinlinefont .}) {\color[rgb]{0,0,0}\plot 14.417 17.185 15.204 16.514 / }% \linethickness= 0.500pt \setplotsymbol ({\thinlinefont .}) {\color[rgb]{0,0,0}\plot 9.131 15.035 10.234 13.851 / }% \linethickness= 0.500pt \setplotsymbol ({\thinlinefont .}) {\color[rgb]{0,0,0}\plot 10.433 14.916 11.024 14.385 / }% \linethickness= 0.500pt \setplotsymbol ({\thinlinefont .}) {\color[rgb]{0,0,0}\plot 11.557 15.035 11.891 14.719 / }% \put{\SetFigFont{6}{7.2}{\rmdefault}{\mddefault}{\updefault}{\color[rgb]{0,0,0}$\Gamma_u$}% } [lB] at 8.350 15.240 \put{\SetFigFont{6}{7.2}{\rmdefault}{\mddefault}{\updefault}{\color[rgb]{0,0,0}$\Gamma_d$}% } [lB] at 9.652 12.764 \put{\SetFigFont{6}{7.2}{\rmdefault}{\mddefault}{\updefault}{\color[rgb]{0,0,0}$\tilde{S}_-$}% } [lB] at 3.842 12.637 \put{\SetFigFont{6}{7.2}{\rmdefault}{\mddefault}{\updefault}{\color[rgb]{0,0,0}$\bar{N}_-$}% } [lB] at 3.873 10.160 \put{\SetFigFont{6}{7.2}{\rmdefault}{\mddefault}{\updefault}{\color[rgb]{0,0,0}$\bar{N}_+$}% } [lB] at 15.335 17.780 \put{\SetFigFont{6}{7.2}{\rmdefault}{\mddefault}{\updefault}{\color[rgb]{0,0,0}$\tilde{S}_+$}% } [lB] at 15.526 15.304 \put{\SetFigFont{6}{7.2}{\rmdefault}{\mddefault}{\updefault}{\color[rgb]{0,0,0}${\mathcal N}$}% } [lB] at 9.620 18.605 \put{\SetFigFont{6}{7.2}{\rmdefault}{\mddefault}{\updefault}{\color[rgb]{0,0,0}${\mathcal P}$}% } [lB] at 7.271 12.478 \put{\SetFigFont{6}{7.2}{\rmdefault}{\mddefault}{\updefault}{\color[rgb]{0,0,0}${\mathcal N}$}% } [lB] at 12.668 12.827 \linethickness=0pt \putrectangle corners at 3.810 21.660 and 15.558 6.280 \endpicture} \qquad \font\thinlinefont=cmr5 \begingroup\makeatletter\ifx\SetFigFont\undefined% \gdef\SetFigFont#1#2#3#4#5{% \reset@font\fontsize{#1}{#2pt}% \fontfamily{#3}\fontseries{#4}\fontshape{#5}% \selectfont}% \fi\endgroup% \mbox{\beginpicture \setcoordinatesystem units <0.50000cm,0.50000cm> \unitlength=0.50000cm \linethickness=1pt \setplotsymbol ({\makebox(0,0)[l]{\tencirc\symbol{'160}}}) \setshadesymbol ({\thinlinefont .}) \setlinear \linethickness= 0.500pt \setplotsymbol ({\thinlinefont .}) {\color[rgb]{0,0,0}\putrule from 5.080 16.510 to 15.240 16.510 }% \linethickness= 0.500pt \setplotsymbol ({\thinlinefont .}) {\color[rgb]{0,0,0}\putrule from 5.080 11.430 to 15.240 11.430 }% \linethickness=1pt \setplotsymbol ({\makebox(0,0)[l]{\tencirc\symbol{'160}}}) {\color[rgb]{0,0,0}\putrule from 5.080 21.590 to 15.240 21.590 }% \linethickness=1pt \setplotsymbol ({\makebox(0,0)[l]{\tencirc\symbol{'160}}}) {\color[rgb]{0,0,0}\putrule from 15.240 21.590 to 15.240 6.315 \putrule from 15.240 6.350 to 5.045 6.350 \putrule from 5.080 6.350 to 5.080 21.590 }% \linethickness= 0.500pt \setplotsymbol ({\thinlinefont .}) {\color[rgb]{0,0,0}\plot 13.970 16.192 14.446 15.875 / \plot 14.200 15.963 14.446 15.875 14.270 16.069 / }% \linethickness= 0.500pt \setplotsymbol ({\thinlinefont .}) {\color[rgb]{0,0,0}\putrule from 5.080 17.462 to 5.080 16.034 \plot 5.017 16.288 5.080 16.034 5.143 16.288 / }% \linethickness= 0.500pt \setplotsymbol ({\thinlinefont .}) {\color[rgb]{0,0,0}\putrule from 15.240 12.383 to 15.240 10.795 \plot 15.176 11.049 15.240 10.795 15.304 11.049 / }% \linethickness= 0.500pt \setplotsymbol ({\thinlinefont .}) \setdashes < 0.1905cm> {\color[rgb]{0,0,0}\plot 5.080 12.700 5.084 12.706 / \plot 5.084 12.706 5.091 12.719 / \plot 5.091 12.719 5.105 12.742 / \plot 5.105 12.742 5.124 12.776 / \plot 5.124 12.776 5.150 12.821 / \plot 5.150 12.821 5.179 12.874 / \plot 5.179 12.874 5.213 12.933 / \plot 5.213 12.933 5.249 12.994 / \plot 5.249 12.994 5.283 13.056 / \plot 5.283 13.056 5.317 13.115 / \plot 5.317 13.115 5.349 13.172 / \plot 5.349 13.172 5.376 13.227 / \plot 5.376 13.227 5.404 13.278 / \plot 5.404 13.278 5.429 13.327 / \plot 5.429 13.327 5.450 13.373 / \plot 5.450 13.373 5.472 13.418 / \plot 5.472 13.418 5.493 13.460 / \plot 5.493 13.460 5.512 13.504 / \plot 5.512 13.504 5.531 13.547 / \plot 5.531 13.547 5.546 13.587 / \plot 5.546 13.587 5.563 13.627 / \plot 5.563 13.627 5.580 13.669 / \plot 5.580 13.669 5.594 13.714 / \plot 5.594 13.714 5.611 13.758 / \plot 5.611 13.758 5.628 13.807 / \plot 5.628 13.807 5.645 13.856 / \plot 5.645 13.856 5.662 13.904 / \plot 5.662 13.904 5.679 13.957 / \plot 5.679 13.957 5.698 14.010 / \plot 5.698 14.010 5.715 14.063 / \plot 5.715 14.063 5.732 14.118 / \plot 5.732 14.118 5.751 14.173 / \plot 5.751 14.173 5.768 14.228 / \plot 5.768 14.228 5.785 14.283 / \plot 5.785 14.283 5.802 14.338 / \plot 5.802 14.338 5.819 14.393 / \plot 5.819 14.393 5.836 14.446 / \plot 5.836 14.446 5.850 14.499 / \plot 5.850 14.499 5.867 14.552 / \plot 5.867 14.552 5.884 14.605 / \plot 5.884 14.605 5.901 14.658 / \plot 5.901 14.658 5.916 14.707 / \plot 5.916 14.707 5.931 14.755 / \plot 5.931 14.755 5.948 14.806 / \plot 5.948 14.806 5.965 14.859 / \plot 5.965 14.859 5.982 14.912 / \plot 5.982 14.912 5.999 14.967 / \plot 5.999 14.967 6.016 15.022 / \plot 6.016 15.022 6.035 15.077 / \plot 6.035 15.077 6.054 15.134 / \plot 6.054 15.134 6.073 15.193 / \plot 6.073 15.193 6.092 15.251 / \plot 6.092 15.251 6.113 15.308 / \plot 6.113 15.308 6.132 15.365 / \plot 6.132 15.365 6.151 15.422 / \plot 6.151 15.422 6.170 15.477 / \plot 6.170 15.477 6.189 15.532 / \plot 6.189 15.532 6.208 15.585 / \plot 6.208 15.585 6.225 15.636 / \plot 6.225 15.636 6.242 15.685 / \plot 6.242 15.685 6.259 15.731 / \plot 6.259 15.731 6.276 15.776 / \plot 6.276 15.776 6.293 15.820 / \plot 6.293 15.820 6.308 15.862 / \plot 6.308 15.862 6.325 15.900 / \plot 6.325 15.900 6.342 15.947 / \plot 6.342 15.947 6.361 15.994 / \plot 6.361 15.994 6.378 16.038 / \plot 6.378 16.038 6.397 16.082 / \plot 6.397 16.082 6.416 16.125 / \plot 6.416 16.125 6.435 16.169 / \plot 6.435 16.169 6.456 16.212 / \plot 6.456 16.212 6.477 16.256 / \plot 6.477 16.256 6.500 16.298 / \plot 6.500 16.298 6.524 16.341 / \plot 6.524 16.341 6.547 16.381 / \plot 6.547 16.381 6.572 16.423 / \plot 6.572 16.423 6.598 16.463 / \plot 6.598 16.463 6.625 16.502 / \plot 6.625 16.502 6.651 16.540 / \plot 6.651 16.540 6.680 16.578 / \plot 6.680 16.578 6.708 16.614 / \plot 6.708 16.614 6.737 16.650 / \plot 6.737 16.650 6.767 16.686 / \plot 6.767 16.686 6.801 16.722 / \plot 6.801 16.722 6.828 16.751 / \plot 6.828 16.751 6.858 16.783 / \plot 6.858 16.783 6.888 16.815 / \plot 6.888 16.815 6.919 16.849 / \plot 6.919 16.849 6.955 16.883 / \plot 6.955 16.883 6.991 16.919 / \plot 6.991 16.919 7.027 16.954 / \plot 7.027 16.954 7.068 16.990 / \plot 7.068 16.990 7.108 17.029 / \plot 7.108 17.029 7.148 17.067 / \plot 7.148 17.067 7.190 17.107 / \plot 7.190 17.107 7.235 17.145 / \plot 7.235 17.145 7.277 17.183 / \plot 7.277 17.183 7.322 17.223 / \plot 7.322 17.223 7.366 17.261 / \plot 7.366 17.261 7.408 17.300 / \plot 7.408 17.300 7.451 17.336 / \plot 7.451 17.336 7.493 17.371 / \plot 7.493 17.371 7.533 17.407 / \plot 7.533 17.407 7.573 17.441 / \plot 7.573 17.441 7.614 17.475 / \plot 7.614 17.475 7.652 17.507 / \plot 7.652 17.507 7.690 17.539 / \plot 7.690 17.539 7.726 17.568 / \plot 7.726 17.568 7.766 17.602 / \plot 7.766 17.602 7.804 17.634 / \plot 7.804 17.634 7.844 17.666 / \plot 7.844 17.666 7.885 17.697 / \plot 7.885 17.697 7.925 17.731 / \plot 7.925 17.731 7.965 17.763 / \plot 7.965 17.763 8.007 17.795 / \plot 8.007 17.795 8.050 17.827 / \plot 8.050 17.827 8.094 17.856 / \plot 8.094 17.856 8.139 17.888 / \plot 8.139 17.888 8.183 17.918 / \plot 8.183 17.918 8.227 17.945 / \plot 8.227 17.945 8.274 17.973 / \plot 8.274 17.973 8.319 17.998 / \plot 8.319 17.998 8.363 18.023 / \plot 8.363 18.023 8.410 18.047 / \plot 8.410 18.047 8.454 18.068 / \plot 8.454 18.068 8.498 18.087 / \plot 8.498 18.087 8.543 18.104 / \plot 8.543 18.104 8.587 18.121 / \plot 8.587 18.121 8.632 18.136 / \plot 8.632 18.136 8.678 18.150 / \plot 8.678 18.150 8.725 18.163 / \plot 8.725 18.163 8.774 18.174 / \plot 8.774 18.174 8.824 18.184 / \plot 8.824 18.184 8.875 18.195 / \plot 8.875 18.195 8.930 18.203 / \plot 8.930 18.203 8.985 18.210 / \plot 8.985 18.210 9.042 18.216 / \plot 9.042 18.216 9.100 18.222 / \plot 9.100 18.222 9.159 18.227 / \plot 9.159 18.227 9.220 18.229 / \plot 9.220 18.229 9.279 18.231 / \plot 9.279 18.231 9.339 18.231 / \plot 9.339 18.231 9.398 18.231 / \plot 9.398 18.231 9.455 18.229 / \plot 9.455 18.229 9.512 18.224 / \plot 9.512 18.224 9.565 18.220 / \plot 9.565 18.220 9.618 18.216 / \plot 9.618 18.216 9.667 18.210 / \plot 9.667 18.210 9.713 18.203 / \plot 9.713 18.203 9.760 18.195 / \plot 9.760 18.195 9.802 18.186 / \plot 9.802 18.186 9.842 18.176 / \plot 9.842 18.176 9.889 18.163 / \plot 9.889 18.163 9.936 18.150 / \plot 9.936 18.150 9.980 18.133 / \plot 9.980 18.133 10.022 18.117 / \plot 10.022 18.117 10.065 18.095 / \plot 10.065 18.095 10.105 18.074 / \plot 10.105 18.074 10.147 18.051 / \plot 10.147 18.051 10.185 18.026 / \plot 10.185 18.026 10.226 18.000 / \plot 10.226 18.000 10.262 17.973 / \plot 10.262 17.973 10.300 17.943 / \plot 10.300 17.943 10.334 17.913 / \plot 10.334 17.913 10.370 17.884 / \plot 10.370 17.884 10.401 17.852 / \plot 10.401 17.852 10.435 17.822 / \plot 10.435 17.822 10.467 17.791 / \plot 10.467 17.791 10.499 17.759 / \plot 10.499 17.759 10.530 17.727 / \plot 10.530 17.727 10.560 17.697 / \plot 10.560 17.697 10.590 17.668 / \plot 10.590 17.668 10.621 17.636 / \plot 10.621 17.636 10.655 17.602 / \plot 10.655 17.602 10.689 17.568 / \plot 10.689 17.568 10.725 17.532 / \plot 10.725 17.532 10.761 17.496 / \plot 10.761 17.496 10.799 17.458 / \plot 10.799 17.458 10.835 17.418 / \plot 10.835 17.418 10.875 17.380 / \plot 10.875 17.380 10.914 17.340 / \plot 10.914 17.340 10.950 17.302 / \plot 10.950 17.302 10.988 17.261 / \plot 10.988 17.261 11.024 17.223 / \plot 11.024 17.223 11.060 17.185 / \plot 11.060 17.185 11.093 17.149 / \plot 11.093 17.149 11.127 17.113 / \plot 11.127 17.113 11.159 17.079 / \plot 11.159 17.079 11.189 17.046 / \plot 11.189 17.046 11.218 17.012 / \plot 11.218 17.012 11.244 16.982 / \plot 11.244 16.982 11.271 16.952 / \plot 11.271 16.952 11.297 16.923 / \plot 11.297 16.923 11.322 16.891 / \plot 11.322 16.891 11.347 16.859 / \plot 11.347 16.859 11.373 16.828 / \plot 11.373 16.828 11.398 16.792 / \plot 11.398 16.792 11.424 16.756 / \plot 11.424 16.756 11.449 16.720 / \plot 11.449 16.720 11.474 16.679 / \plot 11.474 16.679 11.498 16.639 / \plot 11.498 16.639 11.523 16.597 / \plot 11.523 16.597 11.546 16.554 / \plot 11.546 16.554 11.568 16.510 / \plot 11.568 16.510 11.591 16.466 / \plot 11.591 16.466 11.612 16.419 / \plot 11.612 16.419 11.631 16.370 / \plot 11.631 16.370 11.650 16.322 / \plot 11.650 16.322 11.669 16.273 / \plot 11.669 16.273 11.686 16.220 / \plot 11.686 16.220 11.705 16.167 / \plot 11.705 16.167 11.722 16.112 / \plot 11.722 16.112 11.733 16.072 / \plot 11.733 16.072 11.745 16.027 / \plot 11.745 16.027 11.756 15.983 / \plot 11.756 15.983 11.769 15.934 / \plot 11.769 15.934 11.779 15.886 / \plot 11.779 15.886 11.792 15.833 / \plot 11.792 15.833 11.803 15.778 / \plot 11.803 15.778 11.813 15.720 / \plot 11.813 15.720 11.824 15.661 / \plot 11.824 15.661 11.834 15.598 / \plot 11.834 15.598 11.843 15.534 / \plot 11.843 15.534 11.853 15.466 / \plot 11.853 15.466 11.862 15.399 / \plot 11.862 15.399 11.870 15.327 / \plot 11.870 15.327 11.877 15.253 / \plot 11.877 15.253 11.883 15.179 / \plot 11.883 15.179 11.889 15.102 / \plot 11.889 15.102 11.894 15.024 / \plot 11.894 15.024 11.898 14.946 / \plot 11.898 14.946 11.902 14.867 / \plot 11.902 14.867 11.904 14.787 / \plot 11.904 14.787 11.904 14.707 / \plot 11.904 14.707 11.906 14.626 / \plot 11.906 14.626 11.904 14.544 / \plot 11.904 14.544 11.904 14.463 / \plot 11.904 14.463 11.900 14.381 / \plot 11.900 14.381 11.898 14.300 / \plot 11.898 14.300 11.891 14.218 / \plot 11.891 14.218 11.887 14.133 / \plot 11.887 14.133 11.881 14.048 / \plot 11.881 14.048 11.874 13.983 / \plot 11.874 13.983 11.866 13.913 / \plot 11.866 13.913 11.860 13.843 / \plot 11.860 13.843 11.851 13.773 / \plot 11.851 13.773 11.843 13.699 / \plot 11.843 13.699 11.832 13.625 / \plot 11.832 13.625 11.822 13.551 / \plot 11.822 13.551 11.811 13.473 / \plot 11.811 13.473 11.798 13.394 / \plot 11.798 13.394 11.786 13.314 / \plot 11.786 13.314 11.771 13.231 / \plot 11.771 13.231 11.756 13.149 / \plot 11.756 13.149 11.741 13.064 / \plot 11.741 13.064 11.724 12.979 / \plot 11.724 12.979 11.707 12.893 / \plot 11.707 12.893 11.688 12.806 / \plot 11.688 12.806 11.671 12.717 / \plot 11.671 12.717 11.652 12.630 / \plot 11.652 12.630 11.631 12.541 / \plot 11.631 12.541 11.612 12.452 / \plot 11.612 12.452 11.591 12.366 / \plot 11.591 12.366 11.568 12.277 / \plot 11.568 12.277 11.546 12.190 / \plot 11.546 12.190 11.525 12.103 / \plot 11.525 12.103 11.502 12.018 / \plot 11.502 12.018 11.479 11.934 / \plot 11.479 11.934 11.455 11.851 / \plot 11.455 11.851 11.432 11.769 / \plot 11.432 11.769 11.409 11.688 / \plot 11.409 11.688 11.386 11.610 / \plot 11.386 11.610 11.362 11.532 / \plot 11.362 11.532 11.337 11.458 / \plot 11.337 11.458 11.314 11.383 / \plot 11.314 11.383 11.290 11.309 / \plot 11.290 11.309 11.265 11.240 / \plot 11.265 11.240 11.242 11.170 / \plot 11.242 11.170 11.216 11.100 / \plot 11.216 11.100 11.193 11.032 / \plot 11.193 11.032 11.165 10.962 / \plot 11.165 10.962 11.138 10.892 / \plot 11.138 10.892 11.110 10.825 / \plot 11.110 10.825 11.083 10.757 / \plot 11.083 10.757 11.053 10.689 / \plot 11.053 10.689 11.024 10.621 / \plot 11.024 10.621 10.992 10.554 / \plot 10.992 10.554 10.958 10.488 / \plot 10.958 10.488 10.924 10.422 / \plot 10.924 10.422 10.888 10.357 / \plot 10.888 10.357 10.850 10.291 / \plot 10.850 10.291 10.812 10.228 / \plot 10.812 10.228 10.770 10.164 / \plot 10.770 10.164 10.727 10.103 / \plot 10.727 10.103 10.683 10.041 / \plot 10.683 10.041 10.638 9.982 / \plot 10.638 9.982 10.590 9.925 / \plot 10.590 9.925 10.541 9.870 / \plot 10.541 9.870 10.490 9.815 / \plot 10.490 9.815 10.437 9.764 / \plot 10.437 9.764 10.382 9.713 / \plot 10.382 9.713 10.327 9.667 / \plot 10.327 9.667 10.270 9.620 / \plot 10.270 9.620 10.211 9.578 / \plot 10.211 9.578 10.152 9.538 / \plot 10.152 9.538 10.090 9.500 / \plot 10.090 9.500 10.027 9.466 / \plot 10.027 9.466 9.961 9.432 / \plot 9.961 9.432 9.895 9.402 / \plot 9.895 9.402 9.830 9.375 / \plot 9.830 9.375 9.760 9.349 / \plot 9.760 9.349 9.690 9.326 / \plot 9.690 9.326 9.618 9.307 / \plot 9.618 9.307 9.546 9.290 / \plot 9.546 9.290 9.470 9.273 / \plot 9.470 9.273 9.394 9.260 / \plot 9.394 9.260 9.328 9.252 / \plot 9.328 9.252 9.260 9.243 / \plot 9.260 9.243 9.191 9.237 / \plot 9.191 9.237 9.119 9.233 / \plot 9.119 9.233 9.045 9.231 / \plot 9.045 9.231 8.966 9.229 / \plot 8.966 9.229 8.886 9.231 / \plot 8.886 9.231 8.801 9.233 / \plot 8.801 9.233 8.712 9.237 / \plot 8.712 9.237 8.621 9.241 / \plot 8.621 9.241 8.524 9.250 / \plot 8.524 9.250 8.422 9.258 / \plot 8.422 9.258 8.314 9.271 / \plot 8.314 9.271 8.204 9.284 / \plot 8.204 9.284 8.086 9.299 / \plot 8.086 9.299 7.963 9.315 / \plot 7.963 9.315 7.836 9.332 / \plot 7.836 9.332 7.703 9.354 / \plot 7.703 9.354 7.563 9.375 / \plot 7.563 9.375 7.419 9.398 / \plot 7.419 9.398 7.271 9.423 / \plot 7.271 9.423 7.120 9.449 / \plot 7.120 9.449 6.966 9.476 / \plot 6.966 9.476 6.807 9.506 / \plot 6.807 9.506 6.651 9.533 / \plot 6.651 9.533 6.494 9.563 / \plot 6.494 9.563 6.337 9.593 / \plot 6.337 9.593 6.185 9.622 / \plot 6.185 9.622 6.039 9.650 / \plot 6.039 9.650 5.897 9.677 / \plot 5.897 9.677 5.766 9.703 / \plot 5.766 9.703 5.643 9.728 / \plot 5.643 9.728 5.533 9.749 / \plot 5.533 9.749 5.433 9.771 / \plot 5.433 9.771 5.347 9.787 / \plot 5.347 9.787 5.275 9.802 / \plot 5.275 9.802 5.215 9.815 / \plot 5.215 9.815 5.167 9.823 / \plot 5.167 9.823 5.133 9.832 / \plot 5.133 9.832 5.108 9.836 / \plot 5.108 9.836 5.093 9.840 / \plot 5.093 9.840 5.084 9.842 / \plot 5.084 9.842 5.080 9.842 / }% \linethickness= 0.500pt \setplotsymbol ({\thinlinefont .}) {\color[rgb]{0,0,0}\plot 15.240 15.240 15.234 15.244 / \plot 15.234 15.244 15.221 15.255 / \plot 15.221 15.255 15.196 15.274 / \plot 15.196 15.274 15.164 15.301 / \plot 15.164 15.301 15.121 15.335 / \plot 15.121 15.335 15.073 15.373 / \plot 15.073 15.373 15.022 15.416 / \plot 15.022 15.416 14.973 15.456 / \plot 14.973 15.456 14.925 15.498 / \plot 14.925 15.498 14.878 15.536 / \plot 14.878 15.536 14.838 15.574 / \plot 14.838 15.574 14.800 15.608 / \plot 14.800 15.608 14.768 15.642 / \plot 14.768 15.642 14.736 15.674 / \plot 14.736 15.674 14.709 15.706 / \plot 14.709 15.706 14.683 15.737 / \plot 14.683 15.737 14.658 15.769 / \plot 14.658 15.769 14.635 15.803 / \plot 14.635 15.803 14.611 15.837 / \plot 14.611 15.837 14.588 15.873 / \plot 14.588 15.873 14.565 15.909 / \plot 14.565 15.909 14.544 15.949 / \plot 14.544 15.949 14.520 15.989 / \plot 14.520 15.989 14.499 16.032 / \plot 14.499 16.032 14.478 16.076 / \plot 14.478 16.076 14.457 16.121 / \plot 14.457 16.121 14.438 16.165 / \plot 14.438 16.165 14.419 16.212 / \plot 14.419 16.212 14.400 16.256 / \plot 14.400 16.256 14.383 16.300 / \plot 14.383 16.300 14.368 16.345 / \plot 14.368 16.345 14.353 16.387 / \plot 14.353 16.387 14.340 16.430 / \plot 14.340 16.430 14.326 16.470 / \plot 14.326 16.470 14.315 16.510 / \plot 14.315 16.510 14.302 16.550 / \plot 14.302 16.550 14.290 16.593 / \plot 14.290 16.593 14.279 16.633 / \plot 14.279 16.633 14.268 16.677 / \plot 14.268 16.677 14.258 16.720 / \plot 14.258 16.720 14.247 16.766 / \plot 14.247 16.766 14.237 16.811 / \plot 14.237 16.811 14.228 16.857 / \plot 14.228 16.857 14.220 16.904 / \plot 14.220 16.904 14.211 16.950 / \plot 14.211 16.950 14.205 16.995 / \plot 14.205 16.995 14.199 17.039 / \plot 14.199 17.039 14.192 17.081 / \plot 14.192 17.081 14.188 17.124 / \plot 14.188 17.124 14.186 17.164 / \plot 14.186 17.164 14.184 17.202 / \plot 14.184 17.202 14.182 17.240 / \plot 14.182 17.240 14.182 17.276 / \plot 14.182 17.276 14.182 17.314 / \plot 14.182 17.314 14.184 17.350 / \plot 14.184 17.350 14.186 17.386 / \plot 14.186 17.386 14.190 17.422 / \plot 14.190 17.422 14.196 17.460 / \plot 14.196 17.460 14.203 17.496 / \plot 14.203 17.496 14.211 17.534 / \plot 14.211 17.534 14.222 17.570 / \plot 14.222 17.570 14.235 17.606 / \plot 14.235 17.606 14.247 17.642 / \plot 14.247 17.642 14.264 17.674 / \plot 14.264 17.674 14.281 17.708 / \plot 14.281 17.708 14.300 17.738 / \plot 14.300 17.738 14.321 17.765 / \plot 14.321 17.765 14.343 17.793 / \plot 14.343 17.793 14.368 17.816 / \plot 14.368 17.816 14.393 17.839 / \plot 14.393 17.839 14.421 17.858 / \plot 14.421 17.858 14.446 17.877 / \plot 14.446 17.877 14.476 17.894 / \plot 14.476 17.894 14.506 17.909 / \plot 14.506 17.909 14.541 17.924 / \plot 14.541 17.924 14.580 17.939 / \plot 14.580 17.939 14.624 17.954 / \plot 14.624 17.954 14.673 17.968 / \plot 14.673 17.968 14.726 17.983 / \plot 14.726 17.983 14.785 18.000 / \plot 14.785 18.000 14.848 18.015 / \plot 14.848 18.015 14.914 18.030 / \plot 14.914 18.030 14.980 18.045 / \plot 14.980 18.045 15.045 18.057 / \plot 15.045 18.057 15.102 18.070 / \plot 15.102 18.070 15.153 18.081 / \plot 15.153 18.081 15.191 18.089 / \plot 15.191 18.089 15.219 18.093 / \plot 15.219 18.093 15.234 18.095 / \plot 15.234 18.095 15.240 18.098 / }% \linethickness= 0.500pt \setplotsymbol ({\thinlinefont .}) \setsolid {\color[rgb]{0,0,0}\plot 5.080 12.700 5.082 12.706 / \plot 5.082 12.706 5.088 12.717 / \plot 5.088 12.717 5.101 12.740 / \plot 5.101 12.740 5.118 12.774 / \plot 5.118 12.774 5.141 12.819 / \plot 5.141 12.819 5.171 12.876 / \plot 5.171 12.876 5.207 12.941 / \plot 5.207 12.941 5.245 13.013 / \plot 5.245 13.013 5.285 13.089 / \plot 5.285 13.089 5.328 13.168 / \plot 5.328 13.168 5.370 13.244 / \plot 5.370 13.244 5.412 13.320 / \plot 5.412 13.320 5.453 13.392 / \plot 5.453 13.392 5.491 13.460 / \plot 5.491 13.460 5.529 13.523 / \plot 5.529 13.523 5.565 13.583 / \plot 5.565 13.583 5.599 13.638 / \plot 5.599 13.638 5.632 13.691 / \plot 5.632 13.691 5.664 13.739 / \plot 5.664 13.739 5.696 13.786 / \plot 5.696 13.786 5.730 13.830 / \plot 5.730 13.830 5.762 13.875 / \plot 5.762 13.875 5.795 13.917 / \plot 5.795 13.917 5.825 13.955 / \plot 5.825 13.955 5.857 13.995 / \plot 5.857 13.995 5.891 14.034 / \plot 5.891 14.034 5.925 14.074 / \plot 5.925 14.074 5.961 14.112 / \plot 5.961 14.112 5.997 14.152 / \plot 5.997 14.152 6.035 14.192 / \plot 6.035 14.192 6.073 14.232 / \plot 6.073 14.232 6.113 14.275 / \plot 6.113 14.275 6.155 14.315 / \plot 6.155 14.315 6.198 14.357 / \plot 6.198 14.357 6.242 14.398 / \plot 6.242 14.398 6.284 14.440 / \plot 6.284 14.440 6.329 14.480 / \plot 6.329 14.480 6.375 14.520 / \plot 6.375 14.520 6.420 14.561 / \plot 6.420 14.561 6.464 14.601 / \plot 6.464 14.601 6.509 14.639 / \plot 6.509 14.639 6.553 14.677 / \plot 6.553 14.677 6.598 14.713 / \plot 6.598 14.713 6.642 14.751 / \plot 6.642 14.751 6.684 14.785 / \plot 6.684 14.785 6.727 14.821 / \plot 6.727 14.821 6.769 14.855 / \plot 6.769 14.855 6.811 14.889 / \plot 6.811 14.889 6.854 14.922 / \plot 6.854 14.922 6.892 14.954 / \plot 6.892 14.954 6.932 14.986 / \plot 6.932 14.986 6.972 15.018 / \plot 6.972 15.018 7.013 15.050 / \plot 7.013 15.050 7.055 15.081 / \plot 7.055 15.081 7.097 15.115 / \plot 7.097 15.115 7.142 15.149 / \plot 7.142 15.149 7.188 15.183 / \plot 7.188 15.183 7.235 15.219 / \plot 7.235 15.219 7.283 15.255 / \plot 7.283 15.255 7.332 15.291 / \plot 7.332 15.291 7.383 15.329 / \plot 7.383 15.329 7.434 15.365 / \plot 7.434 15.365 7.487 15.403 / \plot 7.487 15.403 7.540 15.439 / \plot 7.540 15.439 7.595 15.477 / \plot 7.595 15.477 7.648 15.515 / \plot 7.648 15.515 7.703 15.551 / \plot 7.703 15.551 7.758 15.589 / \plot 7.758 15.589 7.813 15.625 / \plot 7.813 15.625 7.868 15.661 / \plot 7.868 15.661 7.923 15.697 / \plot 7.923 15.697 7.978 15.731 / \plot 7.978 15.731 8.033 15.765 / \plot 8.033 15.765 8.088 15.801 / \plot 8.088 15.801 8.143 15.835 / \plot 8.143 15.835 8.198 15.869 / \plot 8.198 15.869 8.255 15.900 / \plot 8.255 15.900 8.306 15.932 / \plot 8.306 15.932 8.357 15.962 / \plot 8.357 15.962 8.410 15.991 / \plot 8.410 15.991 8.465 16.021 / \plot 8.465 16.021 8.520 16.053 / \plot 8.520 16.053 8.577 16.085 / \plot 8.577 16.085 8.636 16.116 / \plot 8.636 16.116 8.697 16.148 / \plot 8.697 16.148 8.759 16.180 / \plot 8.759 16.180 8.822 16.212 / \plot 8.822 16.212 8.888 16.245 / \plot 8.888 16.245 8.956 16.277 / \plot 8.956 16.277 9.023 16.309 / \plot 9.023 16.309 9.091 16.341 / \plot 9.091 16.341 9.163 16.375 / \plot 9.163 16.375 9.233 16.404 / \plot 9.233 16.404 9.305 16.436 / \plot 9.305 16.436 9.377 16.466 / \plot 9.377 16.466 9.449 16.495 / \plot 9.449 16.495 9.521 16.525 / \plot 9.521 16.525 9.593 16.552 / \plot 9.593 16.552 9.665 16.578 / \plot 9.665 16.578 9.735 16.603 / \plot 9.735 16.603 9.807 16.626 / \plot 9.807 16.626 9.874 16.650 / \plot 9.874 16.650 9.944 16.671 / \plot 9.944 16.671 10.012 16.692 / \plot 10.012 16.692 10.080 16.711 / \plot 10.080 16.711 10.147 16.728 / \plot 10.147 16.728 10.213 16.745 / \plot 10.213 16.745 10.279 16.760 / \plot 10.279 16.760 10.346 16.775 / \plot 10.346 16.775 10.412 16.787 / \plot 10.412 16.787 10.478 16.800 / \plot 10.478 16.800 10.545 16.811 / \plot 10.545 16.811 10.613 16.821 / \plot 10.613 16.821 10.681 16.830 / \plot 10.681 16.830 10.751 16.836 / \plot 10.751 16.836 10.823 16.842 / \plot 10.823 16.842 10.894 16.849 / \plot 10.894 16.849 10.966 16.851 / \plot 10.966 16.851 11.041 16.855 / \putrule from 11.041 16.855 to 11.115 16.855 \putrule from 11.115 16.855 to 11.191 16.855 \plot 11.191 16.855 11.267 16.853 / \plot 11.267 16.853 11.343 16.849 / \plot 11.343 16.849 11.419 16.844 / \plot 11.419 16.844 11.496 16.838 / \plot 11.496 16.838 11.570 16.830 / \plot 11.570 16.830 11.646 16.819 / \plot 11.646 16.819 11.720 16.808 / \plot 11.720 16.808 11.794 16.796 / \plot 11.794 16.796 11.866 16.781 / \plot 11.866 16.781 11.938 16.766 / \plot 11.938 16.766 12.008 16.749 / \plot 12.008 16.749 12.076 16.730 / \plot 12.076 16.730 12.141 16.711 / \plot 12.141 16.711 12.207 16.690 / \plot 12.207 16.690 12.270 16.667 / \plot 12.270 16.667 12.332 16.643 / \plot 12.332 16.643 12.393 16.618 / \plot 12.393 16.618 12.452 16.593 / \plot 12.452 16.593 12.509 16.565 / \plot 12.509 16.565 12.569 16.535 / \plot 12.569 16.535 12.624 16.506 / \plot 12.624 16.506 12.681 16.474 / \plot 12.681 16.474 12.736 16.440 / \plot 12.736 16.440 12.791 16.404 / \plot 12.791 16.404 12.848 16.366 / \plot 12.848 16.366 12.903 16.328 / \plot 12.903 16.328 12.958 16.286 / \plot 12.958 16.286 13.013 16.241 / \plot 13.013 16.241 13.070 16.195 / \plot 13.070 16.195 13.125 16.148 / \plot 13.125 16.148 13.180 16.097 / \plot 13.180 16.097 13.236 16.044 / \plot 13.236 16.044 13.291 15.991 / \plot 13.291 15.991 13.343 15.936 / \plot 13.343 15.936 13.396 15.879 / \plot 13.396 15.879 13.449 15.820 / \plot 13.449 15.820 13.502 15.761 / \plot 13.502 15.761 13.551 15.699 / \plot 13.551 15.699 13.602 15.638 / \plot 13.602 15.638 13.648 15.574 / \plot 13.648 15.574 13.695 15.513 / \plot 13.695 15.513 13.739 15.450 / \plot 13.739 15.450 13.784 15.386 / \plot 13.784 15.386 13.824 15.323 / \plot 13.824 15.323 13.864 15.259 / \plot 13.864 15.259 13.902 15.196 / \plot 13.902 15.196 13.940 15.132 / \plot 13.940 15.132 13.974 15.069 / \plot 13.974 15.069 14.008 15.007 / \plot 14.008 15.007 14.042 14.944 / \plot 14.042 14.944 14.072 14.880 / \plot 14.072 14.880 14.103 14.817 / \plot 14.103 14.817 14.129 14.760 / \plot 14.129 14.760 14.154 14.702 / \plot 14.154 14.702 14.177 14.643 / \plot 14.177 14.643 14.203 14.584 / \plot 14.203 14.584 14.226 14.522 / \plot 14.226 14.522 14.249 14.461 / \plot 14.249 14.461 14.273 14.398 / \plot 14.273 14.398 14.294 14.332 / \plot 14.294 14.332 14.317 14.264 / \plot 14.317 14.264 14.338 14.196 / \plot 14.338 14.196 14.359 14.127 / \plot 14.359 14.127 14.381 14.055 / \plot 14.381 14.055 14.402 13.981 / \plot 14.402 13.981 14.423 13.904 / \plot 14.423 13.904 14.442 13.826 / \plot 14.442 13.826 14.461 13.748 / \plot 14.461 13.748 14.482 13.667 / \plot 14.482 13.667 14.499 13.587 / \plot 14.499 13.587 14.518 13.504 / \plot 14.518 13.504 14.535 13.420 / \plot 14.535 13.420 14.554 13.335 / \plot 14.554 13.335 14.571 13.250 / \plot 14.571 13.250 14.586 13.164 / \plot 14.586 13.164 14.603 13.079 / \plot 14.603 13.079 14.618 12.992 / \plot 14.618 12.992 14.633 12.903 / \plot 14.633 12.903 14.647 12.816 / \plot 14.647 12.816 14.662 12.730 / \plot 14.662 12.730 14.675 12.641 / \plot 14.675 12.641 14.690 12.552 / \plot 14.690 12.552 14.702 12.463 / \plot 14.702 12.463 14.715 12.374 / \plot 14.715 12.374 14.728 12.285 / \plot 14.728 12.285 14.738 12.196 / \plot 14.738 12.196 14.751 12.105 / \plot 14.751 12.105 14.764 12.012 / \plot 14.764 12.012 14.774 11.936 / \plot 14.774 11.936 14.783 11.858 / \plot 14.783 11.858 14.793 11.779 / \plot 14.793 11.779 14.804 11.697 / \plot 14.804 11.697 14.815 11.614 / \plot 14.815 11.614 14.825 11.529 / \plot 14.825 11.529 14.834 11.441 / \plot 14.834 11.441 14.846 11.350 / \plot 14.846 11.350 14.857 11.256 / \plot 14.857 11.256 14.867 11.157 / \plot 14.867 11.157 14.878 11.055 / \plot 14.878 11.055 14.891 10.950 / \plot 14.891 10.950 14.903 10.839 / \plot 14.903 10.839 14.916 10.723 / \plot 14.916 10.723 14.929 10.605 / \plot 14.929 10.605 14.942 10.480 / \plot 14.942 10.480 14.956 10.348 / \plot 14.956 10.348 14.971 10.215 / \plot 14.971 10.215 14.986 10.075 / \plot 14.986 10.075 15.001 9.931 / \plot 15.001 9.931 15.016 9.783 / \plot 15.016 9.783 15.033 9.631 / \plot 15.033 9.631 15.050 9.478 / \plot 15.050 9.478 15.064 9.322 / \plot 15.064 9.322 15.081 9.165 / \plot 15.081 9.165 15.098 9.011 / \plot 15.098 9.011 15.113 8.856 / \plot 15.113 8.856 15.128 8.706 / \plot 15.128 8.706 15.143 8.560 / \plot 15.143 8.560 15.157 8.422 / \plot 15.157 8.422 15.172 8.293 / \plot 15.172 8.293 15.183 8.172 / \plot 15.183 8.172 15.196 8.064 / \plot 15.196 8.064 15.204 7.967 / \plot 15.204 7.967 15.212 7.882 / \plot 15.212 7.882 15.221 7.811 / \plot 15.221 7.811 15.227 7.751 / \plot 15.227 7.751 15.232 7.705 / \plot 15.232 7.705 15.236 7.671 / \plot 15.236 7.671 15.238 7.648 / \putrule from 15.238 7.648 to 15.238 7.631 \plot 15.238 7.631 15.240 7.624 / \putrule from 15.240 7.624 to 15.240 7.620 }% \linethickness= 0.500pt \setplotsymbol ({\thinlinefont .}) {\color[rgb]{0,0,0}\plot 15.240 15.240 15.236 15.242 / \plot 15.236 15.242 15.229 15.248 / \plot 15.229 15.248 15.215 15.261 / \plot 15.215 15.261 15.193 15.278 / \plot 15.193 15.278 15.164 15.304 / \plot 15.164 15.304 15.124 15.335 / \plot 15.124 15.335 15.077 15.373 / \plot 15.077 15.373 15.024 15.418 / \plot 15.024 15.418 14.963 15.466 / \plot 14.963 15.466 14.897 15.519 / \plot 14.897 15.519 14.829 15.574 / \plot 14.829 15.574 14.760 15.632 / \plot 14.760 15.632 14.688 15.687 / \plot 14.688 15.687 14.616 15.744 / \plot 14.616 15.744 14.546 15.797 / \plot 14.546 15.797 14.476 15.852 / \plot 14.476 15.852 14.408 15.903 / \plot 14.408 15.903 14.343 15.953 / \plot 14.343 15.953 14.275 16.002 / \plot 14.275 16.002 14.211 16.049 / \plot 14.211 16.049 14.146 16.095 / \plot 14.146 16.095 14.080 16.142 / \plot 14.080 16.142 14.014 16.188 / \plot 14.014 16.188 13.947 16.235 / \plot 13.947 16.235 13.877 16.281 / \plot 13.877 16.281 13.805 16.328 / \plot 13.805 16.328 13.733 16.377 / \plot 13.733 16.377 13.678 16.413 / \plot 13.678 16.413 13.623 16.449 / \plot 13.623 16.449 13.568 16.485 / \plot 13.568 16.485 13.509 16.521 / \plot 13.509 16.521 13.447 16.559 / \plot 13.447 16.559 13.386 16.597 / \plot 13.386 16.597 13.322 16.635 / \plot 13.322 16.635 13.255 16.673 / \plot 13.255 16.673 13.187 16.711 / \plot 13.187 16.711 13.115 16.749 / \plot 13.115 16.749 13.041 16.789 / \plot 13.041 16.789 12.967 16.828 / \plot 12.967 16.828 12.888 16.866 / \plot 12.888 16.866 12.808 16.904 / \plot 12.808 16.904 12.725 16.942 / \plot 12.725 16.942 12.641 16.978 / \plot 12.641 16.978 12.556 17.014 / \plot 12.556 17.014 12.467 17.050 / \plot 12.467 17.050 12.378 17.081 / \plot 12.378 17.081 12.287 17.115 / \plot 12.287 17.115 12.194 17.145 / \plot 12.194 17.145 12.101 17.175 / \plot 12.101 17.175 12.006 17.200 / \plot 12.006 17.200 11.910 17.225 / \plot 11.910 17.225 11.815 17.249 / \plot 11.815 17.249 11.718 17.268 / \plot 11.718 17.268 11.621 17.287 / \plot 11.621 17.287 11.523 17.304 / \plot 11.523 17.304 11.426 17.316 / \plot 11.426 17.316 11.328 17.327 / \plot 11.328 17.327 11.229 17.333 / \plot 11.229 17.333 11.132 17.340 / \plot 11.132 17.340 11.034 17.342 / \putrule from 11.034 17.342 to 10.937 17.342 \plot 10.937 17.342 10.837 17.338 / \plot 10.837 17.338 10.740 17.331 / \plot 10.740 17.331 10.643 17.323 / \plot 10.643 17.323 10.543 17.310 / \plot 10.543 17.310 10.446 17.295 / \plot 10.446 17.295 10.346 17.276 / \plot 10.346 17.276 10.253 17.257 / \plot 10.253 17.257 10.162 17.236 / \plot 10.162 17.236 10.069 17.211 / \plot 10.069 17.211 9.974 17.183 / \plot 9.974 17.183 9.878 17.156 / \plot 9.878 17.156 9.781 17.124 / \plot 9.781 17.124 9.682 17.090 / \plot 9.682 17.090 9.582 17.054 / \plot 9.582 17.054 9.481 17.016 / \plot 9.481 17.016 9.379 16.978 / \plot 9.379 16.978 9.273 16.938 / \plot 9.273 16.938 9.169 16.895 / \plot 9.169 16.895 9.064 16.851 / \plot 9.064 16.851 8.956 16.806 / \plot 8.956 16.806 8.850 16.762 / \plot 8.850 16.762 8.740 16.717 / \plot 8.740 16.717 8.632 16.671 / \plot 8.632 16.671 8.524 16.626 / \plot 8.524 16.626 8.414 16.580 / \plot 8.414 16.580 8.306 16.535 / \plot 8.306 16.535 8.198 16.491 / \plot 8.198 16.491 8.090 16.447 / \plot 8.090 16.447 7.984 16.404 / \plot 7.984 16.404 7.878 16.362 / \plot 7.878 16.362 7.775 16.324 / \plot 7.775 16.324 7.671 16.286 / \plot 7.671 16.286 7.569 16.250 / \plot 7.569 16.250 7.470 16.216 / \plot 7.470 16.216 7.374 16.184 / \plot 7.374 16.184 7.279 16.157 / \plot 7.279 16.157 7.186 16.131 / \plot 7.186 16.131 7.095 16.108 / \plot 7.095 16.108 7.008 16.089 / \plot 7.008 16.089 6.924 16.072 / \plot 6.924 16.072 6.841 16.059 / \plot 6.841 16.059 6.761 16.051 / \plot 6.761 16.051 6.682 16.044 / \plot 6.682 16.044 6.608 16.042 / \plot 6.608 16.042 6.538 16.044 / \plot 6.538 16.044 6.469 16.049 / \plot 6.469 16.049 6.403 16.059 / \plot 6.403 16.059 6.337 16.072 / \plot 6.337 16.072 6.276 16.091 / \plot 6.276 16.091 6.219 16.112 / \plot 6.219 16.112 6.164 16.140 / \plot 6.164 16.140 6.111 16.169 / \plot 6.111 16.169 6.060 16.205 / \plot 6.060 16.205 6.013 16.245 / \plot 6.013 16.245 5.967 16.290 / \plot 5.967 16.290 5.922 16.341 / \plot 5.922 16.341 5.880 16.398 / \plot 5.880 16.398 5.840 16.461 / \plot 5.840 16.461 5.800 16.529 / \plot 5.800 16.529 5.762 16.605 / \plot 5.762 16.605 5.726 16.688 / \plot 5.726 16.688 5.690 16.777 / \plot 5.690 16.777 5.656 16.872 / \plot 5.656 16.872 5.622 16.976 / \plot 5.622 16.976 5.590 17.086 / \plot 5.590 17.086 5.558 17.204 / \plot 5.558 17.204 5.529 17.329 / \plot 5.529 17.329 5.499 17.462 / \plot 5.499 17.462 5.469 17.602 / \plot 5.469 17.602 5.442 17.748 / \plot 5.442 17.748 5.412 17.903 / \plot 5.412 17.903 5.387 18.062 / \plot 5.387 18.062 5.359 18.227 / \plot 5.359 18.227 5.334 18.396 / \plot 5.334 18.396 5.309 18.570 / \plot 5.309 18.570 5.285 18.743 / \plot 5.285 18.743 5.262 18.919 / \plot 5.262 18.919 5.241 19.094 / \plot 5.241 19.094 5.220 19.266 / \plot 5.220 19.266 5.201 19.435 / \plot 5.201 19.435 5.182 19.596 / \plot 5.182 19.596 5.165 19.751 / \plot 5.165 19.751 5.150 19.895 / \plot 5.150 19.895 5.137 20.028 / \plot 5.137 20.028 5.124 20.149 / \plot 5.124 20.149 5.114 20.256 / \plot 5.114 20.256 5.105 20.350 / \plot 5.105 20.350 5.099 20.428 / \plot 5.099 20.428 5.093 20.494 / \plot 5.093 20.494 5.088 20.544 / \plot 5.088 20.544 5.084 20.582 / \plot 5.084 20.582 5.082 20.608 / \plot 5.082 20.608 5.080 20.625 / \putrule from 5.080 20.625 to 5.080 20.633 \putrule from 5.080 20.633 to 5.080 20.637 }% \linethickness= 0.500pt \setplotsymbol ({\thinlinefont .}) {\color[rgb]{0,0,0}\plot 5.175 11.430 7.017 9.525 / }% \linethickness= 0.500pt \setplotsymbol ({\thinlinefont .}) {\color[rgb]{0,0,0}\plot 5.175 12.668 8.509 9.271 / }% \linethickness= 0.500pt \setplotsymbol ({\thinlinefont .}) {\color[rgb]{0,0,0}\plot 5.524 13.462 9.716 9.398 / }% \linethickness= 0.500pt \setplotsymbol ({\thinlinefont .}) {\color[rgb]{0,0,0}\plot 5.874 14.415 10.446 9.874 / }% \linethickness= 0.500pt \setplotsymbol ({\thinlinefont .}) {\color[rgb]{0,0,0}\plot 6.159 15.367 10.986 10.700 / }% \linethickness= 0.500pt \setplotsymbol ({\thinlinefont .}) {\color[rgb]{0,0,0}\plot 6.445 16.320 11.335 11.621 / }% \linethickness= 0.500pt \setplotsymbol ({\thinlinefont .}) {\color[rgb]{0,0,0}\plot 7.017 16.986 11.557 12.573 / }% \linethickness= 0.500pt \setplotsymbol ({\thinlinefont .}) {\color[rgb]{0,0,0}\plot 7.747 17.526 11.716 13.589 / }% \linethickness= 0.500pt \setplotsymbol ({\thinlinefont .}) {\color[rgb]{0,0,0}\plot 8.319 18.129 11.811 14.796 / }% \linethickness= 0.500pt \setplotsymbol ({\thinlinefont .}) {\color[rgb]{0,0,0}\plot 9.493 18.193 11.462 16.510 / }% \linethickness= 0.500pt \setplotsymbol ({\thinlinefont .}) {\color[rgb]{0,0,0}\plot 5.143 10.192 5.461 9.811 / }% \linethickness= 0.500pt \setplotsymbol ({\thinlinefont .}) {\color[rgb]{0,0,0}\plot 14.383 16.224 15.145 15.589 / }% \linethickness= 0.500pt \setplotsymbol ({\thinlinefont .}) {\color[rgb]{0,0,0}\plot 14.129 17.558 15.113 16.637 / }% \linethickness= 0.500pt \setplotsymbol ({\thinlinefont .}) {\color[rgb]{0,0,0}\plot 14.859 18.034 15.145 17.780 / }% \put{\SetFigFont{6}{7.2}{\rmdefault}{\mddefault}{\updefault}{\color[rgb]{0,0,0}$\Gamma_l$}% } [lB] at 11.906 13.653 \put{\SetFigFont{6}{7.2}{\rmdefault}{\mddefault}{\updefault}{\color[rgb]{0,0,0}$\Gamma_r$}% } [lB] at 13.494 17.462 \put{\SetFigFont{6}{7.2}{\rmdefault}{\mddefault}{\updefault}{\color[rgb]{0,0,0}${\mathcal P}_l$}% } [lB] at 14.446 16.669 \put{\SetFigFont{6}{7.2}{\rmdefault}{\mddefault}{\updefault}{\color[rgb]{0,0,0}${\mathcal P}_r$}% } [lB] at 6.985 12.859 \put{\SetFigFont{6}{7.2}{\rmdefault}{\mddefault}{\updefault}{\color[rgb]{0,0,0}${\mathcal N}$}% } [lB] at 10.160 19.685 \put{\SetFigFont{6}{7.2}{\rmdefault}{\mddefault}{\updefault}{\color[rgb]{0,0,0}$\tilde{S}_-$}% } [lB] at 3.842 12.668 \put{\SetFigFont{6}{7.2}{\rmdefault}{\mddefault}{\updefault}{\color[rgb]{0,0,0} $\tilde{S}_+$}% } [lB] at 15.367 15.208 \put{\SetFigFont{6}{7.2}{\rmdefault}{\mddefault}{\updefault}{\color[rgb]{0,0,0}$\tilde{N}_+$}% } [lB] at 15.367 18.002 \put{\SetFigFont{6}{7.2}{\rmdefault}{\mddefault}{\updefault}{\color[rgb]{0,0,0}$\bar{N}_+$}% } [lB] at 15.399 7.588 \put{\SetFigFont{6}{7.2}{\rmdefault}{\mddefault}{\updefault}{\color[rgb]{0,0,0}$\bar{N}_-$}% } [lB] at 3.873 9.716 \put{\SetFigFont{6}{7.2}{\rmdefault}{\mddefault}{\updefault}{\color[rgb]{0,0,0}$\tilde{N}_-$}% } [lB] at 3.842 20.225 \put{\SetFigFont{6}{7.2}{\rmdefault}{\mddefault}{\updefault}{\color[rgb]{0,0,0}${\mathcal N}$}% } [lB] at 12.732 14.732 \linethickness=0pt \putrectangle corners at 3.810 21.660 and 15.431 6.280 \endpicture} \end{center} \caption{Change in the topology of nullclines: $\mu<\mu_c$ (left) and $\mu>\mu_c$ (right).\label{fig:topnul}} \end{figure} \subsubsection{Winding number of orbits and corridors} Assuming \refeq{assump}, let \begin{equation} w_0 := \frac{s_+^{}- n_-^{}}{2\pi}= \frac{n_+^{}-s_-^{}}{2\pi}.\end{equation} Given any orbit $\mathcal{O} = \{(x_o(\tau),y_o(\tau))\ | \ \tau\in{\mathbb R}\}$ of the flow \refeq{eq:flow}, the following quantity is well-defined: \begin{equation}\label{def:wn} w(\mathcal{O}) = w_0 - \frac{1}{2\pi} \left(y_o(\infty) - y_o(-\infty)\right) = w_0-\frac{1}{2\pi}\int_{-\infty}^\infty g_\mu(x_o(\tau),y_o(\tau)) d\tau \end{equation} In particular for the two distinguished orbits $\widetilde{\mathcal{W}}^\pm$ this is easily calculated to be \begin{equation} w(\widetilde{\mathcal{W}}^-) = w_0 - \frac{1}{2\pi}(n_+^{} - 2\pi k_+ - s_-^{})= k_+ \qquad w(\widetilde{\mathcal{W}}^+) = w_0-\frac{1}{2\pi}\left(s_+^{} - ( n_-^{} + 2\pi k_-)\right)= k_- \end{equation} and is therefore an integer. We call these the {\em winding numbers}, of $\mathcal{W}^-$ and $\mathcal{W}^+$, respectively. On the other hand it is easy to see that these two must in fact be equal. The reason is that, since $\widetilde{\mathcal{W}}^-$ goes from $\widetilde{S}^- = (x_-^{},s_-^{})$ to $\bar{N}^+=(x_+^{},n_+^{}-2\pi k_+)$, there are two copies of it in the universal cover that go from $(x_-^{},s_-^{} + 2\pi k_+)$ to $(x_+^{},n_+^{})$ and from $(x_-^{},s_-^{}+2\pi(k_+-1))$ to $(x_+^{},n_+^{}-2\pi)$. Since $n_+^{}-2\pi<s_+^{}<n_+^{}$, these two copies of $\widetilde{\mathcal{W}}^-$ must sandwich any orbit that goes into $\widetilde{S}_+$, in particular $\widetilde{\mathcal{W}}^+$. Thus the $\alpha$-limit point of $\widetilde{\mathcal{W}}^+$ has to be $(x_-^{},n_-^{}+2\pi k_+)$, and therefore $k_+=k_-$. \begin{defn} The {\em winding number of the corridor} ${\mathcal K}_1$ is the common value of the winding numbers of $\widetilde{\mathcal{W}}^\pm$. \end{defn} Figure~\ref{fig:topnul} shows a corridor of winding number zero when $\mu<\mu_c$ (left) and one of winding number equal to one for $\mu>\mu_c$ (right). \subsubsection{Continuity argument for existence of saddles connectors} As the parameter $\mu$ varies, the two distinguished orbits $\widetilde{\mathcal{W}}^\pm$, and hence the corridor ${\mathcal K}_1$ that they form will also vary. Let ${\mathcal K}_1(\mu)$ denote the corresponding corridor for parameter value $\mu$ (if it exists). Since the winding number $w({\mathcal K}_1(\mu))$ is integer-valued, if ${\mathcal K}_1$ varies continuously with respect to $\mu$, its winding number would have to remain constant. It is however possible that for some value of $\mu$ the two orbits $\widetilde{\mathcal{W}}^\pm$ coincide and the corridor ${\mathcal K}_1(\mu)$ disappears, leaving a saddles connector behind (for which the winding number will not be an integer). We would like to show that this is the only way for the winding number of ${\mathcal K}_1(\mu)$ to be different for two different values of the parameter $\mu$. In other words: \begin{prop} \label{prop:ssexists} Suppose there exist two values $\mu^{}_0<\mu^{}_1$ in the interval $I$ for each of which a non-empty corridor ${\mathcal K}_1$ of finite winding number exists, and such that \begin{equation} w({\mathcal K}_1(\mu^{}_0))\leq 0\qquad\mbox{and}\qquad w({\mathcal K}_1(\mu^{}_1))\geq 1. \end{equation} Then there exists $\mu^{}_s \in (\mu^{}_0,\mu^{}_1)$ such that the flow \refeq{eq:flowmu} with $\mu=\mu^{}_s$ has a saddles connector ${\mathcal S}(\mu^{}_s)$, whose lift to the universal cover $\bar{\mathcal{C}}$ connects the saddle-node $\widetilde{S}_-$ to the saddle-node $\widetilde{S}_+$. \end{prop} \begin{proof} Let $\mathbf{a}(\mu)$ denote the {\em signed area} of ${\mathcal K}_1(\mu)$, defined via Green's theorem: \begin{equation}\label{def:area} \mathbf{a}(\mu) = \oint_{\partial {\mathcal K}_1(\mu)} (-y)dx = \int_{x_-}^{x_+} \left(y_\mu^+ - y_\mu^-\right) dx. \end{equation} Here $y_\mu^\pm$ are the $y$-components of the orbits $\widetilde{\mathcal{W}}^\pm$, thought of as functions of $x$ (which is always possible since $x(\tau)$ is a monotone increasing function of flow parameter $\tau$.) The following facts about $\mathbf{a}(\mu)$ are easily verified: \begin{enumerate} \item $\mathbf{a}(\mu)>0$ if and only if $w({\mathcal K}_1(\mu))\geq 1$. {\em Proof:} Since $\widetilde{\mathcal{W}}^+$ and $\widetilde{\mathcal{W}}^-$ cannot intersect without coinciding, it is clear from \refeq{def:area} that $\mathbf{a}(\mu)>0$ if and only if $\widetilde{\mathcal{W}}^+$ is above $\widetilde{\mathcal{W}}^-$, which is equivalent to $\widetilde{S}_+ = (x_+,s_+)$ being above $\bar{N}_+ = (x_+,n_+-2\pi k)$ where $k=w({\mathcal K}_1(\mu))$. Since $n_+>s_+$ we must have $k\geq 1$. \item $\mathbf{a}(\mu)<0$ if and only if $w({\mathcal K}_1(\mu)) \leq 0$. {\em Proof:} Similar to above. \item $\mathbf{a}(\mu) = 0$ if and only if ${\mathcal K}_1(\mu) = \emptyset$. {\em Proof:} From the definition \refeq{def:area} it is clear that the only way for $\mathbf{a}(\mu)$ to be zero is for the two orbits $\widetilde{\mathcal{W}}^+$ and $\widetilde{\mathcal{W}}^-$ to coincide. \item $\mathbf{a}(\mu)$ is a continuous function of $\mu$ for all $\mu\in [\mu_0,\mu_1]$. {\em Proof:} Let $\epsilon>0$ be given. For $\mu,\mu'\in [\mu_0,\mu_1]$ and $\xi>0$ small enough, we have \begin{eqnarray*} |\mathbf{a}(\mu') - \mathbf{a}(\mu)| & = &\left| \int_{x_-}^{x_+} \left(y_{\mu'}^+ - y_{\mu}^+\right) - \left( y_{\mu'}^- - y_{\mu}^-\right) dx \right| \\ & \leq & \int_{x_-}^{x_-+\xi} |y_{\mu'}^+| + |y_{\mu}^+| dx + \int_{x_-+\xi}^{x_+-\xi} \left|y_{\mu'}^+ - y_{\mu}^+\right| dx + \int_{x_+-\xi}^{x_+} |y_{\mu'}^+| + |y_{\mu}^+| dx \\ & & \mbox{} + \int_{x_-}^{x_-+\xi} |y_{\mu'}^-| + |y_{\mu}^-| dx + \int_{x_-+\xi}^{x_+ - \xi} \left|y_{\mu'}^- - y_{\mu}^-\right| dx + \int_{x_+-\xi}^{x_+} |y_{\mu'}^-| + |y_{\mu}^-| dx \\ & = &\mbox{I}+\mbox{II}+\mbox{III}+\mbox{IV}+\mbox{V}+\mbox{VI}. \end{eqnarray*} By Lemma~\ref{lem:mon} the functions $y_\mu^\pm$ are monotone in $\mu$, thus for all $\tau \in {\mathbb R}$ we have $$ |y_\mu^\pm(\tau)| \leq \max\{ |y_{\mu_0}^\pm(\tau)|, |y_{\mu_1}^\pm(\tau)|\} $$ Thus is particular, using the continuity of orbits of \refeq{eq:flowmu} in the flow parameter $\tau$, for $x$ near the boundary points $x_\pm$ and all $\mu\in[\mu_0,\mu_1]$ we have $|y_\mu^\pm(x)| \leq C$, where $C>0$ is a constant depending only on $y^\pm_{\mu_0}(x_\pm)$ and $y^\pm_{\mu_1}(x_\pm)$, or in other words, on the finite winding numbers of the corridors ${\mathcal K}_1(\mu_0)$ and ${\mathcal K}_1(\mu_1)$. Thus, given $\epsilon$, we may choose $\xi$ small enough (depending on $\epsilon$) such that $\mbox{I},\mbox{III},\mbox{IV}$ and $\mbox{VI}$ are all less than $\epsilon/6$. Fixing $\xi$ in this way, by continuous dependence of orbits of the flow \refeq{eq:flowmu} on the parameter $\mu$ (and since we are on a compact interval in the flow parameter $\tau$), it is possible for $\delta>0$ to be chosen small enough (depending on $\epsilon$ and $\xi$,) such that $|\mu' - \mu|<\delta$ implies $|y_{\mu'}^+(x) - y_{\mu}^+(x)| < (x_+ - x_-)^{-1}\epsilon /6$ for all $x \in [x_-+\xi,x_+-\xi]$. Therefore $\mbox{II} < \epsilon/6$. Similarly, $\mbox{V}<\epsilon/6$ as well, and we are done. \end{enumerate} Having established the above properties for the signed area $\mathbf{a}(\mu)$, the standard continuity argument can now be applied: By assumption we have $\mathbf{a}(\mu_0)<0$ and $\mathbf{a}(\mu_1)>0$. Thus by the Intermediate Value Theorem there exists $\mu_s\in(\mu_0,\mu_1)$ such that $\mathbf{a}(\mu_s)=0$, which is equivalent to the existence of a connector. Since the corridors ${\mathcal K}_1(\mu)$ by definition always have $\widetilde{S}_\pm$ on their boundary, the saddles connector also has to go from $\widetilde{S}_-$ to $\widetilde{S}_+$. \end{proof} We will use this proposition to establish the existence of saddles connectors for both the $\Theta$ \refeq{dynsysTh} and the $\Omega$ \refeq{dynsysOm} equations. In each case we need to show that the flow satisfies the assumptions we have made in this subsection about the flow on a cylinder \refeq{eq:flowmu}. \subsubsection{Existence of saddles connectors for the $\Theta$ equation} For simplicity we are only going to consider the case $\kappa = \frac{1}{2}$. To see that \refeq{dynsysTh} is a flow on a cylinder of the type we have considered in the above, we make the following identifications: $x=\theta$, $y=\Theta$, $x_-^{} = 0$, $x_+^{} = \pi$. We have $$ f(\theta) = \sin\theta, \qquad g(\theta,\Theta) = -2a\sin\theta\cos\theta\cos\Theta+\left(2aE\sin^2\theta - 1\right) \sin\Theta + 2\lambda\sin\theta. $$ Therefore, the equilibria are at $$ S_- = (0,0),\quad N_- = (0,\pi)\qquad S_+ = (\pi,-\pi),\quad N_+ = (\pi,0); $$ all of these are hyperbolic. Condition \refeq{assump} is clearly satisfied: $s_-^{} - n_-^{} = -\pi$ and $n_+^{} - s_+^{} = \pi$. Here we make a note of the fact, easily verified, that the $\Theta$ flow \refeq{dynsysTh} possesses a discrete symmetry: Let $\mathcal{O} = (\theta(\tau),\Theta(\tau))$ be any orbit of the flow \refeq{dynsysTh}. Then $\mathcal{O}' = (\pi - \theta(-t), \pi-\Theta(-t))$ is also an orbit of \refeq{dynsysTh}. This is simply due to the fact that $$ f(\theta) = f(\pi-\theta),\qquad g(\theta,\Theta) = g(\pi-\theta,\pi-\Theta). $$ This symmetry will prove useful in constructing corridors of given winding number for the flow. Next we check that the assumptions we made in the previous subsection about the topology of nullclines hold in the case of \refeq{dynsysTh}. \subsubsection{Topology of the nullclines} Let $T := \tan(\Theta/2)$. Then, $$ g_{E,\lambda}(\theta,\Theta) = \frac{1}{1+T^2}\left( 2\sin\theta(\lambda - a\cos\theta)T^2 +2(2aE\sin^2\theta - 1)T + 2\sin\theta(\lambda+ a\cos\theta)\right) =: \frac{q(T)}{1+T^2} $$ where $q$ is a quadratic polynomial in $T$, whose discriminant we calculate to be $$ \Delta_q := (2aE\sin^2\theta - 1)^2 - 4\lambda^2\sin^2\theta + 4 a^2 \sin^2\theta\cos^2\theta. $$ Setting $\tau :=- \cot \theta$ we thus obtain $$ \Delta_q = \frac{\delta_q(\tau)}{(1+\tau^2)^2},\qquad \delta_q(\tau) = \tau^4 + 2(1-2aE + 2a^2 - 2\lambda^2)\tau^2+(1-2aE)^2 - 4\lambda^2 $$ Therefore $\delta_q$ is a quartic polynomial which is quadratic in $s=\tau^2$. The discriminant of $\delta_q(s)$ is $$ \Delta_{\delta_q} = \lambda^4 + 2a(E-a)\lambda^2 + a^2(a^2-2aE+1) = (\lambda^2+a(E-a))^2 + a^2(1-E^2) $$ which is always positive. Thus $\delta_q(s) = 0$ will always have two roots, which will be of opposite signs if $|\lambda|\geq\frac{1}{2} - aE$. In that case there exists $s_1 = \tau_1^2>0$ such that $\Delta_q(\pm \tau_1^{})=0$, and $\Delta_q(\tau)<0$ for $\tau\in(-\tau_1^{},\tau_1^{})$. It follows that the quadratic equation $q(T)=0$ will have two roots so long as $|t|>\tau_1^{}$, it will have repeated roots when $t=\pm \tau_1^{}$, and no roots when $|t|<\tau_1^{}$. If on the other hand $|\lambda|<\frac{1}{2}-aE\leq \frac{1}{2}$ then $s_1$ and $s_2$, the roots of $\delta_q(s)=0$, will be of the same sign. Since the sum of these roots will be $4(\lambda^2 - a^2 - (\frac{1}{2}-aE))$ it is easy to see that $s_1$ and $s_2$ will both be negative, thus they will not correspond to real values of $\tau = \pm\sqrt{s}$. Therefore $\Delta_q>0$ and hence $q(T)=0$ will always have two roots. Thus the critical value of the parameter $\lambda$ (thinking of the other parameters $a$ and $E$ as given and fixed) is $$ \lambda_c := -\frac{1}{2} + aE. $$ Now any zero of $q$ will be a zero of $g$, and thus will give us a point on the nullcline $\Gamma$. In addition, $g$ may also have a zero at $\Theta = 0$ or $\pm\pi$, where $T=\pm\infty$. For $g$ to be zero there we need the coefficient of $T^2$ in $q$ to vanish, i.e. either $\sin\theta = 0$, which will give us the equilibrium points, or $\cos\theta = \lambda/a$ which is impossible so long as $$ \lambda < -a. $$ Under this condition therefore, the nullclines have the topology we assumed in the previous subsection, with $\lambda$ playing the role of the parameter $\mu$, i.e., given $a\in(0,\frac{1}{2})$, $E\in[0,1]$, $\lambda\in(-\infty,-a)$ and $\lambda_c$ as in the above, the topology of $\Theta$-nullclines for the flow \refeq{dynsysTh} changes across $\lambda=\lambda_c$ in the manner described in assumption ({\bf A}). Figure~\ref{fig:thetanull} shows Maple plots of the $\Theta$-nullclines for values of $\lambda$ below and above the critical value. \begin{figure}[ht] \begin{center} \includegraphics[scale=0.3]{KTZ_zGKN_I_thetanull1} \includegraphics[scale=0.3]{KTZ_zGKN_I_thetanull2} \end{center} \caption{$\Theta$ nullclines for $\lambda=-0.4$ (left) and $\lambda=-0.9$ (right), with $a=0.1$, $E=0.95$.\label{fig:thetanull}} \end{figure} \subsubsection{Explicit solutions of the $\Theta$ equation}\label{explicit} One easily verifies that given $a\in[0,\frac{1}{2})$, there is an explicit solution of \refeq{eq:Theta} (with $\kappa = \frac{1}{2}$), for $E=1$ and $\lambda = -1+a$, namely $$ \Theta_0(\theta) = - \theta. $$ This furthermore generates a saddles connector for \refeq{dynsysTh}: ${\mathcal S}_0 := (\theta_0(\tau),\Theta_0(\theta_0(\tau))$ where $\theta_0(\tau)$ is the unique solution to the ODE $\dot{\theta} = \sin\theta$ with $\theta(-\infty) = 0$ and $\theta(\infty) = \pi$. This solution will help us get the iteration started. \subsubsection{Existence of corridors with unequal winding number} Throughout this section, $a$ will be a fixed number in $(0,\frac{1}{2})$. The following two propositions will start things off: \begin{prop}\label{prop:lacorlow} Given $E\in[0,1]$ and $\lambda \leq \lambda_l := -1-a$ the flow \refeq{dynsysTh} has a corridor ${\mathcal K}_1(E,\lambda)$ with $w({\mathcal K}_1)\geq 1$. \end{prop} \begin{proof} The linearization of the flow at $S_-$ gives us eigenvalues $\lambda_1= 1$, $\lambda_2=-1$ and a corresponding set of eigenvectors is $v_1 = \left(\begin{array}{c} 1\\ \lambda-a\end{array}\right)$ and $v_2 = \left(\begin{array}{c} 0\\ 1\end{array}\right)$. The orbit $\mathcal{W}^-$, being the unstable manifold of $S_-$ is tangent to the unstable direction $v_1$, therefore the slope of $\mathcal{W}^-$ at $S_-$ is $\frac{d\Theta}{d\theta}|_{\mathcal{W}^-}(S_-) = \lambda - a < 0$. The slope of $\mathcal{W}^+$ at $S_+$ is similarly calculated, and it turns out to be the same $\frac{d\Theta}{d\theta}|_{\mathcal{W}^+}(S_+) = \lambda - a$. \begin{figure}[ht] \begin{center} \font\thinlinefont=cmr5 \begingroup\makeatletter\ifx\SetFigFont\undefined% \gdef\SetFigFont#1#2#3#4#5{% \reset@font\fontsize{#1}{#2pt}% \fontfamily{#3}\fontseries{#4}\fontshape{#5}% \selectfont}% \fi\endgroup% \mbox{\beginpicture \setcoordinatesystem units <0.50000cm,0.50000cm> \unitlength=0.50000cm \linethickness=1pt \setplotsymbol ({\makebox(0,0)[l]{\tencirc\symbol{'160}}}) \setshadesymbol ({\thinlinefont .}) \setlinear \linethickness= 0.500pt \setplotsymbol ({\thinlinefont .}) {\color[rgb]{0,0,0}\putrule from 5.080 16.510 to 15.240 16.510 }% \linethickness=1pt \setplotsymbol ({\makebox(0,0)[l]{\tencirc\symbol{'160}}}) {\color[rgb]{0,0,0}\putrule from 5.080 21.590 to 15.240 21.590 }% \linethickness=1pt \setplotsymbol ({\makebox(0,0)[l]{\tencirc\symbol{'160}}}) {\color[rgb]{0,0,0}\putrule from 15.240 21.590 to 15.240 6.315 \putrule from 15.240 6.350 to 5.045 6.350 \putrule from 5.080 6.350 to 5.080 21.590 }% \linethickness= 0.500pt \setplotsymbol ({\thinlinefont .}) {\color[rgb]{0,0,0}\putrule from 5.080 11.430 to 15.240 11.430 }% \linethickness= 0.500pt \setplotsymbol ({\thinlinefont .}) {\color[rgb]{0,0,0}\plot 5.080 16.510 15.240 11.430 / }% \linethickness= 0.500pt \setplotsymbol ({\thinlinefont .}) {\color[rgb]{0,0,0}\plot 8.096 15.240 8.572 14.446 / \plot 8.387 14.631 8.572 14.446 8.496 14.697 / }% \linethickness= 0.500pt \setplotsymbol ({\thinlinefont .}) {\color[rgb]{0,0,0}\plot 11.748 13.335 12.224 12.700 / \plot 12.021 12.865 12.224 12.700 12.122 12.941 / }% \linethickness= 0.500pt \setplotsymbol ({\thinlinefont .}) {\color[rgb]{0,0,0}\plot 5.080 16.510 5.082 16.506 / \plot 5.082 16.506 5.086 16.493 / \plot 5.086 16.493 5.093 16.474 / \plot 5.093 16.474 5.103 16.442 / \plot 5.103 16.442 5.118 16.398 / \plot 5.118 16.398 5.137 16.343 / \plot 5.137 16.343 5.160 16.277 / \plot 5.160 16.277 5.186 16.201 / \plot 5.186 16.201 5.215 16.116 / \plot 5.215 16.116 5.245 16.025 / \plot 5.245 16.025 5.277 15.932 / \plot 5.277 15.932 5.309 15.839 / \plot 5.309 15.839 5.338 15.744 / \plot 5.338 15.744 5.368 15.653 / \plot 5.368 15.653 5.395 15.564 / \plot 5.395 15.564 5.423 15.477 / \plot 5.423 15.477 5.448 15.397 / \plot 5.448 15.397 5.472 15.318 / \plot 5.472 15.318 5.493 15.244 / \plot 5.493 15.244 5.512 15.174 / \plot 5.512 15.174 5.529 15.107 / \plot 5.529 15.107 5.546 15.041 / \plot 5.546 15.041 5.560 14.978 / \plot 5.560 14.978 5.573 14.916 / \plot 5.573 14.916 5.586 14.857 / \plot 5.586 14.857 5.599 14.798 / \plot 5.599 14.798 5.609 14.736 / \plot 5.609 14.736 5.620 14.677 / \plot 5.620 14.677 5.628 14.618 / \plot 5.628 14.618 5.637 14.556 / \plot 5.637 14.556 5.645 14.493 / \plot 5.645 14.493 5.654 14.431 / \plot 5.654 14.431 5.660 14.366 / \plot 5.660 14.366 5.666 14.300 / \plot 5.666 14.300 5.671 14.235 / \plot 5.671 14.235 5.675 14.167 / \plot 5.675 14.167 5.679 14.097 / \plot 5.679 14.097 5.681 14.029 / \plot 5.681 14.029 5.683 13.959 / \plot 5.683 13.959 5.685 13.890 / \putrule from 5.685 13.890 to 5.685 13.818 \putrule from 5.685 13.818 to 5.685 13.748 \plot 5.685 13.748 5.683 13.680 / \plot 5.683 13.680 5.681 13.610 / \plot 5.681 13.610 5.679 13.542 / \plot 5.679 13.542 5.675 13.477 / \plot 5.675 13.477 5.671 13.411 / \plot 5.671 13.411 5.666 13.346 / \plot 5.666 13.346 5.660 13.284 / \plot 5.660 13.284 5.654 13.223 / \plot 5.654 13.223 5.645 13.164 / \plot 5.645 13.164 5.637 13.106 / \plot 5.637 13.106 5.628 13.049 / \plot 5.628 13.049 5.620 12.992 / \plot 5.620 12.992 5.609 12.937 / \plot 5.609 12.937 5.599 12.880 / \plot 5.599 12.880 5.586 12.821 / \plot 5.586 12.821 5.571 12.764 / \plot 5.571 12.764 5.556 12.704 / \plot 5.556 12.704 5.539 12.645 / \plot 5.539 12.645 5.520 12.581 / \plot 5.520 12.581 5.501 12.518 / \plot 5.501 12.518 5.480 12.450 / \plot 5.480 12.450 5.455 12.380 / \plot 5.455 12.380 5.429 12.306 / \plot 5.429 12.306 5.400 12.230 / \plot 5.400 12.230 5.370 12.150 / \plot 5.370 12.150 5.338 12.067 / \plot 5.338 12.067 5.304 11.982 / \plot 5.304 11.982 5.273 11.898 / \plot 5.273 11.898 5.239 11.813 / \plot 5.239 11.813 5.207 11.735 / \plot 5.207 11.735 5.175 11.661 / \plot 5.175 11.661 5.150 11.597 / \plot 5.150 11.597 5.127 11.542 / \plot 5.127 11.542 5.110 11.498 / \plot 5.110 11.498 5.095 11.468 / \plot 5.095 11.468 5.086 11.447 / \plot 5.086 11.447 5.082 11.434 / \plot 5.082 11.434 5.080 11.430 / }% \linethickness= 0.500pt \setplotsymbol ({\thinlinefont .}) {\color[rgb]{0,0,0}\plot 15.240 16.510 15.238 16.506 / \plot 15.238 16.506 15.234 16.493 / \plot 15.234 16.493 15.227 16.474 / \plot 15.227 16.474 15.217 16.442 / \plot 15.217 16.442 15.202 16.398 / \plot 15.202 16.398 15.183 16.343 / \plot 15.183 16.343 15.160 16.275 / \plot 15.160 16.275 15.134 16.199 / \plot 15.134 16.199 15.105 16.114 / \plot 15.105 16.114 15.075 16.023 / \plot 15.075 16.023 15.043 15.930 / \plot 15.043 15.930 15.011 15.835 / \plot 15.011 15.835 14.982 15.740 / \plot 14.982 15.740 14.952 15.646 / \plot 14.952 15.646 14.925 15.558 / \plot 14.925 15.558 14.897 15.471 / \plot 14.897 15.471 14.872 15.388 / \plot 14.872 15.388 14.848 15.308 / \plot 14.848 15.308 14.827 15.234 / \plot 14.827 15.234 14.808 15.162 / \plot 14.808 15.162 14.791 15.092 / \plot 14.791 15.092 14.774 15.026 / \plot 14.774 15.026 14.760 14.961 / \plot 14.760 14.961 14.747 14.897 / \plot 14.747 14.897 14.734 14.836 / \plot 14.734 14.836 14.721 14.772 / \plot 14.721 14.772 14.711 14.711 / \plot 14.711 14.711 14.700 14.647 / \plot 14.700 14.647 14.692 14.586 / \plot 14.692 14.586 14.683 14.520 / \plot 14.683 14.520 14.675 14.455 / \plot 14.675 14.455 14.666 14.389 / \plot 14.666 14.389 14.660 14.321 / \plot 14.660 14.321 14.654 14.252 / \plot 14.654 14.252 14.649 14.182 / \plot 14.649 14.182 14.645 14.110 / \plot 14.645 14.110 14.641 14.038 / \plot 14.641 14.038 14.639 13.964 / \plot 14.639 13.964 14.637 13.892 / \plot 14.637 13.892 14.635 13.818 / \putrule from 14.635 13.818 to 14.635 13.744 \putrule from 14.635 13.744 to 14.635 13.669 \plot 14.635 13.669 14.637 13.595 / \plot 14.637 13.595 14.639 13.523 / \plot 14.639 13.523 14.641 13.454 / \plot 14.641 13.454 14.645 13.384 / \plot 14.645 13.384 14.649 13.316 / \plot 14.649 13.316 14.654 13.248 / \plot 14.654 13.248 14.660 13.185 / \plot 14.660 13.185 14.666 13.121 / \plot 14.666 13.121 14.675 13.060 / \plot 14.675 13.060 14.683 13.001 / \plot 14.683 13.001 14.692 12.943 / \plot 14.692 12.943 14.700 12.888 / \plot 14.700 12.888 14.711 12.831 / \plot 14.711 12.831 14.721 12.774 / \plot 14.721 12.774 14.734 12.717 / \plot 14.734 12.717 14.749 12.660 / \plot 14.749 12.660 14.764 12.603 / \plot 14.764 12.603 14.781 12.543 / \plot 14.781 12.543 14.800 12.486 / \plot 14.800 12.486 14.819 12.425 / \plot 14.819 12.425 14.840 12.361 / \plot 14.840 12.361 14.865 12.296 / \plot 14.865 12.296 14.891 12.228 / \plot 14.891 12.228 14.920 12.156 / \plot 14.920 12.156 14.950 12.082 / \plot 14.950 12.082 14.982 12.006 / \plot 14.982 12.006 15.016 11.930 / \plot 15.016 11.930 15.047 11.851 / \plot 15.047 11.851 15.081 11.777 / \plot 15.081 11.777 15.113 11.705 / \plot 15.113 11.705 15.145 11.637 / \plot 15.145 11.637 15.170 11.580 / \plot 15.170 11.580 15.193 11.532 / \plot 15.193 11.532 15.210 11.491 / \plot 15.210 11.491 15.225 11.464 / \plot 15.225 11.464 15.234 11.445 / \plot 15.234 11.445 15.238 11.434 / \plot 15.238 11.434 15.240 11.430 / }% \linethickness= 0.500pt \setplotsymbol ({\thinlinefont .}) {\color[rgb]{0,0,0}\plot 5.080 16.510 5.082 16.508 / \plot 5.082 16.508 5.084 16.504 / \plot 5.084 16.504 5.091 16.497 / \plot 5.091 16.497 5.101 16.485 / \plot 5.101 16.485 5.116 16.466 / \plot 5.116 16.466 5.135 16.440 / \plot 5.135 16.440 5.160 16.408 / \plot 5.160 16.408 5.192 16.368 / \plot 5.192 16.368 5.230 16.317 / \plot 5.230 16.317 5.277 16.260 / \plot 5.277 16.260 5.330 16.190 / \plot 5.330 16.190 5.391 16.114 / \plot 5.391 16.114 5.461 16.025 / \plot 5.461 16.025 5.537 15.928 / \plot 5.537 15.928 5.624 15.820 / \plot 5.624 15.820 5.715 15.701 / \plot 5.715 15.701 5.817 15.574 / \plot 5.817 15.574 5.922 15.439 / \plot 5.922 15.439 6.037 15.293 / \plot 6.037 15.293 6.157 15.141 / \plot 6.157 15.141 6.282 14.982 / \plot 6.282 14.982 6.413 14.815 / \plot 6.413 14.815 6.549 14.643 / \plot 6.549 14.643 6.687 14.467 / \plot 6.687 14.467 6.830 14.285 / \plot 6.830 14.285 6.974 14.101 / \plot 6.974 14.101 7.123 13.915 / \plot 7.123 13.915 7.271 13.727 / \plot 7.271 13.727 7.421 13.536 / \plot 7.421 13.536 7.569 13.348 / \plot 7.569 13.348 7.719 13.157 / \plot 7.719 13.157 7.870 12.969 / \plot 7.870 12.969 8.016 12.780 / \plot 8.016 12.780 8.164 12.596 / \plot 8.164 12.596 8.308 12.412 / \plot 8.308 12.412 8.450 12.232 / \plot 8.450 12.232 8.589 12.057 / \plot 8.589 12.057 8.727 11.883 / \plot 8.727 11.883 8.862 11.714 / \plot 8.862 11.714 8.994 11.546 / \plot 8.994 11.546 9.121 11.386 / \plot 9.121 11.386 9.248 11.227 / \plot 9.248 11.227 9.368 11.074 / \plot 9.368 11.074 9.487 10.924 / \plot 9.487 10.924 9.603 10.780 / \plot 9.603 10.780 9.716 10.638 / \plot 9.716 10.638 9.823 10.503 / \plot 9.823 10.503 9.929 10.370 / \plot 9.929 10.370 10.031 10.240 / \plot 10.031 10.240 10.130 10.118 / \plot 10.130 10.118 10.228 9.997 / \plot 10.228 9.997 10.321 9.881 / \plot 10.321 9.881 10.412 9.768 / \plot 10.412 9.768 10.499 9.658 / \plot 10.499 9.658 10.583 9.553 / \plot 10.583 9.553 10.668 9.451 / \plot 10.668 9.451 10.746 9.351 / \plot 10.746 9.351 10.825 9.256 / \plot 10.825 9.256 10.901 9.163 / \plot 10.901 9.163 10.975 9.072 / \plot 10.975 9.072 11.047 8.983 / \plot 11.047 8.983 11.117 8.898 / \plot 11.117 8.898 11.184 8.816 / \plot 11.184 8.816 11.250 8.735 / \plot 11.250 8.735 11.316 8.657 / \plot 11.316 8.657 11.379 8.581 / \plot 11.379 8.581 11.441 8.505 / \plot 11.441 8.505 11.502 8.433 / \plot 11.502 8.433 11.563 8.361 / \plot 11.563 8.361 11.665 8.240 / \plot 11.665 8.240 11.764 8.122 / \plot 11.764 8.122 11.860 8.009 / \plot 11.860 8.009 11.955 7.899 / \plot 11.955 7.899 12.046 7.794 / \plot 12.046 7.794 12.135 7.690 / \plot 12.135 7.690 12.224 7.590 / \plot 12.224 7.590 12.308 7.493 / \plot 12.308 7.493 12.391 7.400 / \plot 12.391 7.400 12.474 7.309 / \plot 12.474 7.309 12.554 7.222 / \plot 12.554 7.222 12.630 7.137 / \plot 12.630 7.137 12.706 7.055 / \plot 12.706 7.055 12.780 6.977 / \plot 12.780 6.977 12.852 6.900 / \plot 12.852 6.900 12.922 6.828 / \plot 12.922 6.828 12.990 6.759 / \plot 12.990 6.759 13.056 6.693 / \plot 13.056 6.693 13.119 6.629 / \plot 13.119 6.629 13.180 6.570 / \plot 13.180 6.570 13.240 6.513 / \plot 13.240 6.513 13.297 6.460 / \plot 13.297 6.460 13.352 6.409 / \plot 13.352 6.409 13.405 6.361 / \plot 13.405 6.361 13.456 6.316 / \plot 13.456 6.316 13.502 6.276 / \plot 13.502 6.276 13.549 6.236 / \plot 13.549 6.236 13.593 6.200 / \plot 13.593 6.200 13.636 6.166 / \plot 13.636 6.166 13.676 6.136 / \plot 13.676 6.136 13.714 6.107 / \plot 13.714 6.107 13.750 6.079 / \plot 13.750 6.079 13.786 6.056 / \plot 13.786 6.056 13.820 6.032 / \plot 13.820 6.032 13.851 6.011 / \plot 13.851 6.011 13.881 5.992 / \plot 13.881 5.992 13.913 5.973 / \plot 13.913 5.973 13.940 5.958 / \plot 13.940 5.958 13.968 5.941 / \plot 13.968 5.941 13.998 5.927 / \plot 13.998 5.927 14.046 5.901 / \plot 14.046 5.901 14.093 5.878 / \plot 14.093 5.878 14.141 5.857 / \plot 14.141 5.857 14.188 5.838 / \plot 14.188 5.838 14.232 5.821 / \plot 14.232 5.821 14.279 5.806 / \plot 14.279 5.806 14.323 5.791 / \plot 14.323 5.791 14.368 5.781 / \plot 14.368 5.781 14.412 5.772 / \plot 14.412 5.772 14.455 5.764 / \plot 14.455 5.764 14.497 5.759 / \plot 14.497 5.759 14.537 5.755 / \putrule from 14.537 5.755 to 14.575 5.755 \putrule from 14.575 5.755 to 14.611 5.755 \plot 14.611 5.755 14.647 5.759 / \plot 14.647 5.759 14.679 5.764 / \plot 14.679 5.764 14.711 5.770 / \plot 14.711 5.770 14.740 5.776 / \plot 14.740 5.776 14.768 5.787 / \plot 14.768 5.787 14.796 5.798 / \plot 14.796 5.798 14.819 5.808 / \plot 14.819 5.808 14.844 5.821 / \plot 14.844 5.821 14.874 5.840 / \plot 14.874 5.840 14.903 5.863 / \plot 14.903 5.863 14.933 5.889 / \plot 14.933 5.889 14.963 5.918 / \plot 14.963 5.918 14.992 5.954 / \plot 14.992 5.954 15.022 5.994 / \plot 15.022 5.994 15.054 6.041 / \plot 15.054 6.041 15.088 6.092 / \plot 15.088 6.092 15.121 6.145 / \plot 15.121 6.145 15.155 6.200 / \plot 15.155 6.200 15.183 6.251 / \plot 15.183 6.251 15.208 6.293 / \plot 15.208 6.293 15.225 6.322 / \plot 15.225 6.322 15.236 6.342 / \plot 15.236 6.342 15.240 6.348 / \putrule from 15.240 6.348 to 15.240 6.350 }% \linethickness= 0.500pt \setplotsymbol ({\thinlinefont .}) {\color[rgb]{0,0,0}\plot 15.240 11.430 15.238 11.432 / \plot 15.238 11.432 15.236 11.436 / \plot 15.236 11.436 15.229 11.445 / \plot 15.229 11.445 15.219 11.458 / \plot 15.219 11.458 15.206 11.479 / \plot 15.206 11.479 15.187 11.504 / \plot 15.187 11.504 15.162 11.540 / \plot 15.162 11.540 15.130 11.582 / \plot 15.130 11.582 15.094 11.635 / \plot 15.094 11.635 15.050 11.697 / \plot 15.050 11.697 14.997 11.769 / \plot 14.997 11.769 14.937 11.853 / \plot 14.937 11.853 14.872 11.946 / \plot 14.872 11.946 14.796 12.050 / \plot 14.796 12.050 14.715 12.164 / \plot 14.715 12.164 14.626 12.287 / \plot 14.626 12.287 14.531 12.421 / \plot 14.531 12.421 14.427 12.565 / \plot 14.427 12.565 14.319 12.715 / \plot 14.319 12.715 14.205 12.874 / \plot 14.205 12.874 14.086 13.039 / \plot 14.086 13.039 13.964 13.212 / \plot 13.964 13.212 13.835 13.390 / \plot 13.835 13.390 13.703 13.572 / \plot 13.703 13.572 13.570 13.758 / \plot 13.570 13.758 13.434 13.947 / \plot 13.434 13.947 13.297 14.137 / \plot 13.297 14.137 13.159 14.330 / \plot 13.159 14.330 13.020 14.522 / \plot 13.020 14.522 12.882 14.715 / \plot 12.882 14.715 12.744 14.906 / \plot 12.744 14.906 12.607 15.096 / \plot 12.607 15.096 12.471 15.284 / \plot 12.471 15.284 12.338 15.469 / \plot 12.338 15.469 12.205 15.653 / \plot 12.205 15.653 12.076 15.831 / \plot 12.076 15.831 11.949 16.006 / \plot 11.949 16.006 11.824 16.180 / \plot 11.824 16.180 11.701 16.347 / \plot 11.701 16.347 11.582 16.510 / \plot 11.582 16.510 11.466 16.669 / \plot 11.466 16.669 11.354 16.825 / \plot 11.354 16.825 11.244 16.976 / \plot 11.244 16.976 11.138 17.122 / \plot 11.138 17.122 11.034 17.264 / \plot 11.034 17.264 10.933 17.399 / \plot 10.933 17.399 10.835 17.532 / \plot 10.835 17.532 10.740 17.661 / \plot 10.740 17.661 10.649 17.786 / \plot 10.649 17.786 10.560 17.907 / \plot 10.560 17.907 10.473 18.023 / \plot 10.473 18.023 10.389 18.136 / \plot 10.389 18.136 10.308 18.246 / \plot 10.308 18.246 10.230 18.351 / \plot 10.230 18.351 10.152 18.453 / \plot 10.152 18.453 10.077 18.553 / \plot 10.077 18.553 10.005 18.648 / \plot 10.005 18.648 9.934 18.741 / \plot 9.934 18.741 9.866 18.832 / \plot 9.866 18.832 9.798 18.919 / \plot 9.798 18.919 9.732 19.006 / \plot 9.732 19.006 9.669 19.088 / \plot 9.669 19.088 9.605 19.169 / \plot 9.605 19.169 9.544 19.247 / \plot 9.544 19.247 9.485 19.325 / \plot 9.485 19.325 9.426 19.399 / \plot 9.426 19.399 9.366 19.473 / \plot 9.366 19.473 9.269 19.594 / \plot 9.269 19.594 9.174 19.713 / \plot 9.174 19.713 9.083 19.827 / \plot 9.083 19.827 8.992 19.937 / \plot 8.992 19.937 8.903 20.043 / \plot 8.903 20.043 8.816 20.146 / \plot 8.816 20.146 8.729 20.248 / \plot 8.729 20.248 8.644 20.345 / \plot 8.644 20.345 8.562 20.441 / \plot 8.562 20.441 8.481 20.534 / \plot 8.481 20.534 8.401 20.623 / \plot 8.401 20.623 8.323 20.709 / \plot 8.323 20.709 8.244 20.792 / \plot 8.244 20.792 8.168 20.872 / \plot 8.168 20.872 8.094 20.951 / \plot 8.094 20.951 8.020 21.025 / \plot 8.020 21.025 7.948 21.097 / \plot 7.948 21.097 7.878 21.167 / \plot 7.878 21.167 7.811 21.232 / \plot 7.811 21.232 7.745 21.294 / \plot 7.745 21.294 7.679 21.353 / \plot 7.679 21.353 7.618 21.410 / \plot 7.618 21.410 7.557 21.463 / \plot 7.557 21.463 7.497 21.514 / \plot 7.497 21.514 7.440 21.560 / \plot 7.440 21.560 7.385 21.605 / \plot 7.385 21.605 7.330 21.647 / \plot 7.330 21.647 7.279 21.685 / \plot 7.279 21.685 7.228 21.721 / \plot 7.228 21.721 7.182 21.755 / \plot 7.182 21.755 7.135 21.787 / \plot 7.135 21.787 7.089 21.816 / \plot 7.089 21.816 7.046 21.844 / \plot 7.046 21.844 7.004 21.867 / \plot 7.004 21.867 6.964 21.891 / \plot 6.964 21.891 6.924 21.914 / \plot 6.924 21.914 6.886 21.933 / \plot 6.886 21.933 6.847 21.952 / \plot 6.847 21.952 6.809 21.971 / \plot 6.809 21.971 6.773 21.986 / \plot 6.773 21.986 6.714 22.013 / \plot 6.714 22.013 6.653 22.037 / \plot 6.653 22.037 6.593 22.060 / \plot 6.593 22.060 6.534 22.081 / \plot 6.534 22.081 6.475 22.098 / \plot 6.475 22.098 6.416 22.115 / \plot 6.416 22.115 6.356 22.130 / \plot 6.356 22.130 6.299 22.145 / \plot 6.299 22.145 6.240 22.155 / \plot 6.240 22.155 6.183 22.164 / \plot 6.183 22.164 6.126 22.172 / \plot 6.126 22.172 6.068 22.176 / \plot 6.068 22.176 6.016 22.181 / \plot 6.016 22.181 5.963 22.183 / \putrule from 5.963 22.183 to 5.912 22.183 \plot 5.912 22.183 5.863 22.181 / \plot 5.863 22.181 5.817 22.178 / \plot 5.817 22.178 5.772 22.174 / \plot 5.772 22.174 5.732 22.168 / \plot 5.732 22.168 5.692 22.159 / \plot 5.692 22.159 5.656 22.151 / \plot 5.656 22.151 5.620 22.142 / \plot 5.620 22.142 5.588 22.130 / \plot 5.588 22.130 5.556 22.119 / \plot 5.556 22.119 5.518 22.102 / \plot 5.518 22.102 5.480 22.083 / \plot 5.480 22.083 5.446 22.060 / \plot 5.446 22.060 5.412 22.035 / \plot 5.412 22.035 5.378 22.007 / \plot 5.378 22.007 5.347 21.973 / \plot 5.347 21.973 5.313 21.935 / \plot 5.313 21.935 5.279 21.895 / \plot 5.279 21.895 5.245 21.848 / \plot 5.245 21.848 5.211 21.800 / \plot 5.211 21.800 5.179 21.753 / \plot 5.179 21.753 5.150 21.706 / \plot 5.150 21.706 5.124 21.666 / \plot 5.124 21.666 5.105 21.632 / \plot 5.105 21.632 5.091 21.609 / \plot 5.091 21.609 5.084 21.596 / \plot 5.084 21.596 5.080 21.590 / }% \put{\SetFigFont{6}{7.2}{\rmdefault}{\mddefault}{\updefault}{\color[rgb]{0,0,0}$\tilde{S}_-$}% } [lB] at 3.810 16.510 \put{\SetFigFont{6}{7.2}{\rmdefault}{\mddefault}{\updefault}{\color[rgb]{0,0,0}$\bar{N}_-$}% } [lB] at 3.810 11.430 \put{\SetFigFont{6}{7.2}{\rmdefault}{\mddefault}{\updefault}{\color[rgb]{0,0,0}$\tilde{N}_-$}% } [lB] at 3.810 21.590 \put{\SetFigFont{6}{7.2}{\rmdefault}{\mddefault}{\updefault}{\color[rgb]{0,0,0}$\tilde{S}_+$}% } [lB] at 15.558 11.430 \put{\SetFigFont{6}{7.2}{\rmdefault}{\mddefault}{\updefault}{\color[rgb]{0,0,0}$\tilde{N}_+$}% } [lB] at 15.558 16.510 \put{\SetFigFont{6}{7.2}{\rmdefault}{\mddefault}{\updefault}{\color[rgb]{0,0,0}$\bar{N}_+$}% } [lB] at 15.558 6.350 \put{\SetFigFont{6}{7.2}{\rmdefault}{\mddefault}{\updefault}{\color[rgb]{0,0,0}$\mathcal{W}^-$}% } [lB] at 10.478 9.842 \put{\SetFigFont{6}{7.2}{\rmdefault}{\mddefault}{\updefault}{\color[rgb]{0,0,0}$\mathcal{W}^+$}% } [lB] at 9.842 19.050 \put{\SetFigFont{6}{7.2}{\rmdefault}{\mddefault}{\updefault}{\color[rgb]{0,0,0}$\Gamma_l$}% } [lB] at 5.715 12.700 \put{\SetFigFont{6}{7.2}{\rmdefault}{\mddefault}{\updefault}{\color[rgb]{0,0,0}$\Gamma_r$}% } [lB] at 14.287 14.605 \linethickness=0pt \putrectangle corners at 3.778 22.229 and 15.589 5.709 \endpicture} \end{center} \caption{Existence of corridor with $w({\mathcal K}_1)\geq 1$} \end{figure} On the other hand, since $\lambda_l<\lambda_c$, we have $\Gamma = \Gamma_l \cup \Gamma_r$. $S_-$ is a terminal point of the curve $\Gamma_l$, on which $g_{E,\lambda}(\theta,\Theta)=0$, hence the slope of $\Gamma_l$ at $S_-$ can be calculated from implicit function theorem to be $$ \left.\frac{d\Theta}{d\theta}\right|_{\Gamma_l}(S_-) = -\frac{\partial_\theta g_{E,\lambda}}{\partial_\Theta g_{E,\lambda}} (S_-) = 2(\lambda - a) < \lambda -a. $$ Same is true for the slope of $\Gamma_r$ at $S_+$, as can be easily verified. Thus $\mathcal{W}^-$ starts off inside ${\mathcal N}$, which is connected, and its initial slope is $\lambda - a < -1$ for $\lambda<-1+a$. Consider the diagonal line segment $S_-S_+$, on which $\Theta = -\theta$. Let us compute the slope of orbits that cross this line, and compare it to the slope of the line: $$ \frac{g_{E,\lambda}(\theta,-\theta)}{f(\theta)} - (-1) = -2a\sin\theta( \cos\theta+E\sin\theta) + 2\lambda +2 \leq 0 \quad\mbox{ for } \lambda \leq -1-a. $$ Thus $S_-S_+$ acts as a ``barrier", not allowing $\mathcal{W}^-$ to cross it from below to above. Hence the $\omega$-limit of $\mathcal{W}^-$ cannot be $(\pi,0)$. The first possible terminal point is then $(\pi,-2\pi)$, hence $w(W^-) \geq 1$. \end{proof} \begin{prop}\label{prop:thlowercor} Suppose that for some $E\in(0,1]$ and some $\lambda\leq \lambda_0 := -1+a$, the flow \refeq{dynsysTh} has a saddles connector $\mathcal{S}^\Theta(E,\lambda)$ whose lift to the universal cover of the cylinder goes from $\widetilde{S}_- = (0,0)$ to $\widetilde{S}_+ = (\pi,-\pi)$. Then, for all $E'\in [0,E)$, there exists a corridor ${\mathcal K}_1(E',\lambda)$ of winding number $w({\mathcal K}_1) = 0$ for \refeq{dynsysTh}. \end{prop} \begin{proof} Let ${\mathcal S}(E,\lambda)= (\theta(\tau),\Theta_{E,\lambda}(\tau))$. Since ${\mathcal S}^\Theta = \mathcal{W}^- = \mathcal{W}^+$, by the calculation done in the proof the previous Proposition, the graph of $\Theta_{E,\lambda}$ is entirely contained in ${\mathcal N}$, thus it has to be monotone decreasing, and thus $\sin\Theta_{E,\lambda}(\tau)\leq 0$ for all $\tau$. \begin{figure}[ht] \begin{center} \font\thinlinefont=cmr5 \begingroup\makeatletter\ifx\SetFigFont\undefined% \gdef\SetFigFont#1#2#3#4#5{% \reset@font\fontsize{#1}{#2pt}% \fontfamily{#3}\fontseries{#4}\fontshape{#5}% \selectfont}% \fi\endgroup% \mbox{\beginpicture \setcoordinatesystem units <0.50000cm,0.50000cm> \unitlength=0.50000cm \linethickness=1pt \setplotsymbol ({\makebox(0,0)[l]{\tencirc\symbol{'160}}}) \setshadesymbol ({\thinlinefont .}) \setlinear \linethickness= 0.500pt \setplotsymbol ({\thinlinefont .}) {\color[rgb]{0,0,0}\putrule from 5.080 16.510 to 15.240 16.510 }% \linethickness=1pt \setplotsymbol ({\makebox(0,0)[l]{\tencirc\symbol{'160}}}) {\color[rgb]{0,0,0}\putrule from 5.080 21.590 to 15.240 21.590 }% \linethickness=1pt \setplotsymbol ({\makebox(0,0)[l]{\tencirc\symbol{'160}}}) {\color[rgb]{0,0,0}\putrule from 15.240 21.590 to 15.240 6.315 \putrule from 15.240 6.350 to 5.045 6.350 \putrule from 5.080 6.350 to 5.080 21.590 }% \linethickness= 0.500pt \setplotsymbol ({\thinlinefont .}) {\color[rgb]{0,0,0}\putrule from 5.080 11.430 to 15.240 11.430 }% \linethickness= 0.500pt \setplotsymbol ({\thinlinefont .}) {\color[rgb]{0,0,0}\plot 7.938 14.287 9.684 13.970 / \plot 9.422 13.953 9.684 13.970 9.445 14.078 / }% \linethickness= 0.500pt \setplotsymbol ({\thinlinefont .}) {\color[rgb]{0,0,0}\plot 11.906 13.018 13.335 12.859 / \plot 13.076 12.824 13.335 12.859 13.090 12.950 / }% \linethickness= 0.500pt \setplotsymbol ({\thinlinefont .}) {\color[rgb]{0,0,0}\plot 7.620 16.828 9.366 16.192 / \plot 9.106 16.220 9.366 16.192 9.149 16.339 / }% \linethickness= 0.500pt \setplotsymbol ({\thinlinefont .}) {\color[rgb]{0,0,0}\plot 12.224 16.828 13.652 16.192 / \plot 13.395 16.238 13.652 16.192 13.446 16.354 / }% \linethickness= 0.500pt \setplotsymbol ({\thinlinefont .}) {\color[rgb]{0,0,0}\putrule from 15.240 12.700 to 15.240 13.970 \plot 15.304 13.716 15.240 13.970 15.176 13.716 / }% \linethickness= 0.500pt \setplotsymbol ({\thinlinefont .}) {\color[rgb]{0,0,0}\plot 5.715 16.510 7.620 14.605 / }% \linethickness= 0.500pt \setplotsymbol ({\thinlinefont .}) {\color[rgb]{0,0,0}\plot 6.985 16.510 9.684 13.811 / }% \linethickness= 0.500pt \setplotsymbol ({\thinlinefont .}) {\color[rgb]{0,0,0}\plot 8.255 16.510 11.430 13.335 / }% \linethickness= 0.500pt \setplotsymbol ({\thinlinefont .}) {\color[rgb]{0,0,0}\plot 9.525 16.510 13.494 12.541 / }% \linethickness= 0.500pt \setplotsymbol ({\thinlinefont .}) {\color[rgb]{0,0,0}\plot 10.795 16.510 15.240 12.065 / }% \linethickness= 0.500pt \setplotsymbol ({\thinlinefont .}) {\color[rgb]{0,0,0}\plot 12.065 16.510 15.240 13.335 / }% \linethickness= 0.500pt \setplotsymbol ({\thinlinefont .}) {\color[rgb]{0,0,0}\plot 13.335 16.510 15.240 14.605 / }% \linethickness= 0.500pt \setplotsymbol ({\thinlinefont .}) {\color[rgb]{0,0,0}\plot 14.605 16.510 15.240 15.875 / }% \linethickness= 0.500pt \setplotsymbol ({\thinlinefont .}) {\color[rgb]{0,0,0}\plot 5.080 16.510 5.082 16.506 / \plot 5.082 16.506 5.086 16.493 / \plot 5.086 16.493 5.093 16.474 / \plot 5.093 16.474 5.103 16.442 / \plot 5.103 16.442 5.118 16.398 / \plot 5.118 16.398 5.137 16.343 / \plot 5.137 16.343 5.160 16.277 / \plot 5.160 16.277 5.186 16.201 / \plot 5.186 16.201 5.215 16.116 / \plot 5.215 16.116 5.245 16.025 / \plot 5.245 16.025 5.277 15.932 / \plot 5.277 15.932 5.309 15.839 / \plot 5.309 15.839 5.338 15.744 / \plot 5.338 15.744 5.368 15.653 / \plot 5.368 15.653 5.395 15.564 / \plot 5.395 15.564 5.423 15.477 / \plot 5.423 15.477 5.448 15.397 / \plot 5.448 15.397 5.472 15.318 / \plot 5.472 15.318 5.493 15.244 / \plot 5.493 15.244 5.512 15.174 / \plot 5.512 15.174 5.529 15.107 / \plot 5.529 15.107 5.546 15.041 / \plot 5.546 15.041 5.560 14.978 / \plot 5.560 14.978 5.573 14.916 / \plot 5.573 14.916 5.586 14.857 / \plot 5.586 14.857 5.599 14.798 / \plot 5.599 14.798 5.609 14.736 / \plot 5.609 14.736 5.620 14.677 / \plot 5.620 14.677 5.628 14.618 / \plot 5.628 14.618 5.637 14.556 / \plot 5.637 14.556 5.645 14.493 / \plot 5.645 14.493 5.654 14.431 / \plot 5.654 14.431 5.660 14.366 / \plot 5.660 14.366 5.666 14.300 / \plot 5.666 14.300 5.671 14.235 / \plot 5.671 14.235 5.675 14.167 / \plot 5.675 14.167 5.679 14.097 / \plot 5.679 14.097 5.681 14.029 / \plot 5.681 14.029 5.683 13.959 / \plot 5.683 13.959 5.685 13.890 / \putrule from 5.685 13.890 to 5.685 13.818 \putrule from 5.685 13.818 to 5.685 13.748 \plot 5.685 13.748 5.683 13.680 / \plot 5.683 13.680 5.681 13.610 / \plot 5.681 13.610 5.679 13.542 / \plot 5.679 13.542 5.675 13.477 / \plot 5.675 13.477 5.671 13.411 / \plot 5.671 13.411 5.666 13.346 / \plot 5.666 13.346 5.660 13.284 / \plot 5.660 13.284 5.654 13.223 / \plot 5.654 13.223 5.645 13.164 / \plot 5.645 13.164 5.637 13.106 / \plot 5.637 13.106 5.628 13.049 / \plot 5.628 13.049 5.620 12.992 / \plot 5.620 12.992 5.609 12.937 / \plot 5.609 12.937 5.599 12.880 / \plot 5.599 12.880 5.586 12.821 / \plot 5.586 12.821 5.571 12.764 / \plot 5.571 12.764 5.556 12.704 / \plot 5.556 12.704 5.539 12.645 / \plot 5.539 12.645 5.520 12.581 / \plot 5.520 12.581 5.501 12.518 / \plot 5.501 12.518 5.480 12.450 / \plot 5.480 12.450 5.455 12.380 / \plot 5.455 12.380 5.429 12.306 / \plot 5.429 12.306 5.400 12.230 / \plot 5.400 12.230 5.370 12.150 / \plot 5.370 12.150 5.338 12.067 / \plot 5.338 12.067 5.304 11.982 / \plot 5.304 11.982 5.273 11.898 / \plot 5.273 11.898 5.239 11.813 / \plot 5.239 11.813 5.207 11.735 / \plot 5.207 11.735 5.175 11.661 / \plot 5.175 11.661 5.150 11.597 / \plot 5.150 11.597 5.127 11.542 / \plot 5.127 11.542 5.110 11.498 / \plot 5.110 11.498 5.095 11.468 / \plot 5.095 11.468 5.086 11.447 / \plot 5.086 11.447 5.082 11.434 / \plot 5.082 11.434 5.080 11.430 / }% \linethickness= 0.500pt \setplotsymbol ({\thinlinefont .}) {\color[rgb]{0,0,0}\plot 15.240 16.510 15.238 16.506 / \plot 15.238 16.506 15.234 16.493 / \plot 15.234 16.493 15.227 16.474 / \plot 15.227 16.474 15.217 16.442 / \plot 15.217 16.442 15.202 16.398 / \plot 15.202 16.398 15.183 16.343 / \plot 15.183 16.343 15.160 16.275 / \plot 15.160 16.275 15.134 16.199 / \plot 15.134 16.199 15.105 16.114 / \plot 15.105 16.114 15.075 16.023 / \plot 15.075 16.023 15.043 15.930 / \plot 15.043 15.930 15.011 15.835 / \plot 15.011 15.835 14.982 15.740 / \plot 14.982 15.740 14.952 15.646 / \plot 14.952 15.646 14.925 15.558 / \plot 14.925 15.558 14.897 15.471 / \plot 14.897 15.471 14.872 15.388 / \plot 14.872 15.388 14.848 15.308 / \plot 14.848 15.308 14.827 15.234 / \plot 14.827 15.234 14.808 15.162 / \plot 14.808 15.162 14.791 15.092 / \plot 14.791 15.092 14.774 15.026 / \plot 14.774 15.026 14.760 14.961 / \plot 14.760 14.961 14.747 14.897 / \plot 14.747 14.897 14.734 14.836 / \plot 14.734 14.836 14.721 14.772 / \plot 14.721 14.772 14.711 14.711 / \plot 14.711 14.711 14.700 14.647 / \plot 14.700 14.647 14.692 14.586 / \plot 14.692 14.586 14.683 14.520 / \plot 14.683 14.520 14.675 14.455 / \plot 14.675 14.455 14.666 14.389 / \plot 14.666 14.389 14.660 14.321 / \plot 14.660 14.321 14.654 14.252 / \plot 14.654 14.252 14.649 14.182 / \plot 14.649 14.182 14.645 14.110 / \plot 14.645 14.110 14.641 14.038 / \plot 14.641 14.038 14.639 13.964 / \plot 14.639 13.964 14.637 13.892 / \plot 14.637 13.892 14.635 13.818 / \putrule from 14.635 13.818 to 14.635 13.744 \putrule from 14.635 13.744 to 14.635 13.669 \plot 14.635 13.669 14.637 13.595 / \plot 14.637 13.595 14.639 13.523 / \plot 14.639 13.523 14.641 13.454 / \plot 14.641 13.454 14.645 13.384 / \plot 14.645 13.384 14.649 13.316 / \plot 14.649 13.316 14.654 13.248 / \plot 14.654 13.248 14.660 13.185 / \plot 14.660 13.185 14.666 13.121 / \plot 14.666 13.121 14.675 13.060 / \plot 14.675 13.060 14.683 13.001 / \plot 14.683 13.001 14.692 12.943 / \plot 14.692 12.943 14.700 12.888 / \plot 14.700 12.888 14.711 12.831 / \plot 14.711 12.831 14.721 12.774 / \plot 14.721 12.774 14.734 12.717 / \plot 14.734 12.717 14.749 12.660 / \plot 14.749 12.660 14.764 12.603 / \plot 14.764 12.603 14.781 12.543 / \plot 14.781 12.543 14.800 12.486 / \plot 14.800 12.486 14.819 12.425 / \plot 14.819 12.425 14.840 12.361 / \plot 14.840 12.361 14.865 12.296 / \plot 14.865 12.296 14.891 12.228 / \plot 14.891 12.228 14.920 12.156 / \plot 14.920 12.156 14.950 12.082 / \plot 14.950 12.082 14.982 12.006 / \plot 14.982 12.006 15.016 11.930 / \plot 15.016 11.930 15.047 11.851 / \plot 15.047 11.851 15.081 11.777 / \plot 15.081 11.777 15.113 11.705 / \plot 15.113 11.705 15.145 11.637 / \plot 15.145 11.637 15.170 11.580 / \plot 15.170 11.580 15.193 11.532 / \plot 15.193 11.532 15.210 11.491 / \plot 15.210 11.491 15.225 11.464 / \plot 15.225 11.464 15.234 11.445 / \plot 15.234 11.445 15.238 11.434 / \plot 15.238 11.434 15.240 11.430 / }% \linethickness= 0.500pt \setplotsymbol ({\thinlinefont .}) {\color[rgb]{0,0,0}\plot 5.080 16.510 5.082 16.508 / \plot 5.082 16.508 5.086 16.504 / \plot 5.086 16.504 5.095 16.493 / \plot 5.095 16.493 5.108 16.478 / \plot 5.108 16.478 5.127 16.457 / \plot 5.127 16.457 5.150 16.430 / \plot 5.150 16.430 5.182 16.396 / \plot 5.182 16.396 5.220 16.353 / \plot 5.220 16.353 5.264 16.305 / \plot 5.264 16.305 5.315 16.248 / \plot 5.315 16.248 5.372 16.188 / \plot 5.372 16.188 5.433 16.121 / \plot 5.433 16.121 5.499 16.053 / \plot 5.499 16.053 5.571 15.979 / \plot 5.571 15.979 5.643 15.905 / \plot 5.643 15.905 5.719 15.831 / \plot 5.719 15.831 5.798 15.754 / \plot 5.798 15.754 5.876 15.680 / \plot 5.876 15.680 5.956 15.606 / \plot 5.956 15.606 6.035 15.534 / \plot 6.035 15.534 6.115 15.464 / \plot 6.115 15.464 6.195 15.397 / \plot 6.195 15.397 6.274 15.331 / \plot 6.274 15.331 6.354 15.268 / \plot 6.354 15.268 6.435 15.208 / \plot 6.435 15.208 6.515 15.149 / \plot 6.515 15.149 6.596 15.094 / \plot 6.596 15.094 6.678 15.039 / \plot 6.678 15.039 6.763 14.986 / \plot 6.763 14.986 6.847 14.935 / \plot 6.847 14.935 6.934 14.887 / \plot 6.934 14.887 7.023 14.838 / \plot 7.023 14.838 7.114 14.791 / \plot 7.114 14.791 7.209 14.743 / \plot 7.209 14.743 7.307 14.696 / \plot 7.307 14.696 7.408 14.652 / \plot 7.408 14.652 7.514 14.605 / \plot 7.514 14.605 7.597 14.569 / \plot 7.597 14.569 7.681 14.535 / \plot 7.681 14.535 7.768 14.499 / \plot 7.768 14.499 7.859 14.465 / \plot 7.859 14.465 7.950 14.429 / \plot 7.950 14.429 8.045 14.393 / \plot 8.045 14.393 8.141 14.357 / \plot 8.141 14.357 8.240 14.321 / \plot 8.240 14.321 8.342 14.285 / \plot 8.342 14.285 8.448 14.247 / \plot 8.448 14.247 8.553 14.211 / \plot 8.553 14.211 8.664 14.173 / \plot 8.664 14.173 8.774 14.135 / \plot 8.774 14.135 8.888 14.097 / \plot 8.888 14.097 9.002 14.059 / \plot 9.002 14.059 9.121 14.021 / \plot 9.121 14.021 9.239 13.981 / \plot 9.239 13.981 9.360 13.942 / \plot 9.360 13.942 9.483 13.904 / \plot 9.483 13.904 9.605 13.864 / \plot 9.605 13.864 9.730 13.824 / \plot 9.730 13.824 9.857 13.786 / \plot 9.857 13.786 9.982 13.746 / \plot 9.982 13.746 10.109 13.705 / \plot 10.109 13.705 10.236 13.667 / \plot 10.236 13.667 10.363 13.627 / \plot 10.363 13.627 10.488 13.589 / \plot 10.488 13.589 10.615 13.551 / \plot 10.615 13.551 10.740 13.513 / \plot 10.740 13.513 10.865 13.473 / \plot 10.865 13.473 10.988 13.437 / \plot 10.988 13.437 11.108 13.399 / \plot 11.108 13.399 11.229 13.360 / \plot 11.229 13.360 11.347 13.324 / \plot 11.347 13.324 11.466 13.288 / \plot 11.466 13.288 11.580 13.252 / \plot 11.580 13.252 11.692 13.219 / \plot 11.692 13.219 11.803 13.183 / \plot 11.803 13.183 11.910 13.149 / \plot 11.910 13.149 12.016 13.115 / \plot 12.016 13.115 12.120 13.083 / \plot 12.120 13.083 12.222 13.049 / \plot 12.222 13.049 12.319 13.018 / \plot 12.319 13.018 12.416 12.986 / \plot 12.416 12.986 12.509 12.956 / \plot 12.509 12.956 12.601 12.924 / \plot 12.601 12.924 12.687 12.895 / \plot 12.687 12.895 12.774 12.865 / \plot 12.774 12.865 12.857 12.835 / \plot 12.857 12.835 12.937 12.806 / \putrule from 12.937 12.806 to 12.939 12.806 \plot 12.939 12.806 13.053 12.764 / \plot 13.053 12.764 13.164 12.721 / \plot 13.164 12.721 13.269 12.679 / \plot 13.269 12.679 13.371 12.636 / \plot 13.371 12.636 13.470 12.594 / \plot 13.470 12.594 13.566 12.552 / \plot 13.566 12.552 13.657 12.509 / \plot 13.657 12.509 13.746 12.465 / \plot 13.746 12.465 13.835 12.421 / \plot 13.835 12.421 13.921 12.374 / \plot 13.921 12.374 14.006 12.325 / \plot 14.006 12.325 14.091 12.275 / \plot 14.091 12.275 14.175 12.224 / \plot 14.175 12.224 14.260 12.171 / \plot 14.260 12.171 14.343 12.116 / \plot 14.343 12.116 14.425 12.059 / \plot 14.425 12.059 14.510 12.002 / \plot 14.510 12.002 14.590 11.944 / \plot 14.590 11.944 14.671 11.885 / \plot 14.671 11.885 14.747 11.826 / \plot 14.747 11.826 14.823 11.769 / \plot 14.823 11.769 14.893 11.716 / \plot 14.893 11.716 14.958 11.663 / \plot 14.958 11.663 15.018 11.616 / \plot 15.018 11.616 15.071 11.572 / \plot 15.071 11.572 15.115 11.536 / \plot 15.115 11.536 15.153 11.504 / \plot 15.153 11.504 15.183 11.479 / \plot 15.183 11.479 15.206 11.460 / \plot 15.206 11.460 15.223 11.445 / \plot 15.223 11.445 15.232 11.436 / \plot 15.232 11.436 15.238 11.432 / \plot 15.238 11.432 15.240 11.430 / }% \linethickness= 0.500pt \setplotsymbol ({\thinlinefont .}) {\color[rgb]{0,0,0}\plot 5.080 16.510 5.084 16.506 / \plot 5.084 16.506 5.091 16.499 / \plot 5.091 16.499 5.105 16.487 / \plot 5.105 16.487 5.127 16.466 / \plot 5.127 16.466 5.156 16.438 / \plot 5.156 16.438 5.194 16.402 / \plot 5.194 16.402 5.241 16.360 / \plot 5.241 16.360 5.292 16.313 / \plot 5.292 16.313 5.347 16.260 / \plot 5.347 16.260 5.406 16.207 / \plot 5.406 16.207 5.465 16.154 / \plot 5.465 16.154 5.527 16.099 / \plot 5.527 16.099 5.588 16.046 / \plot 5.588 16.046 5.645 15.996 / \plot 5.645 15.996 5.702 15.949 / \plot 5.702 15.949 5.759 15.903 / \plot 5.759 15.903 5.812 15.860 / \plot 5.812 15.860 5.863 15.820 / \plot 5.863 15.820 5.914 15.782 / \plot 5.914 15.782 5.965 15.746 / \plot 5.965 15.746 6.013 15.712 / \plot 6.013 15.712 6.064 15.678 / \plot 6.064 15.678 6.115 15.646 / \plot 6.115 15.646 6.166 15.615 / \plot 6.166 15.615 6.219 15.583 / \plot 6.219 15.583 6.259 15.560 / \plot 6.259 15.560 6.301 15.536 / \plot 6.301 15.536 6.344 15.513 / \plot 6.344 15.513 6.390 15.490 / \plot 6.390 15.490 6.437 15.466 / \plot 6.437 15.466 6.485 15.443 / \plot 6.485 15.443 6.536 15.418 / \plot 6.536 15.418 6.591 15.395 / \plot 6.591 15.395 6.646 15.369 / \plot 6.646 15.369 6.703 15.344 / \plot 6.703 15.344 6.765 15.318 / \plot 6.765 15.318 6.828 15.295 / \plot 6.828 15.295 6.894 15.270 / \plot 6.894 15.270 6.964 15.244 / \plot 6.964 15.244 7.036 15.219 / \plot 7.036 15.219 7.110 15.193 / \plot 7.110 15.193 7.186 15.168 / \plot 7.186 15.168 7.264 15.143 / \plot 7.264 15.143 7.345 15.117 / \plot 7.345 15.117 7.429 15.094 / \plot 7.429 15.094 7.514 15.069 / \plot 7.514 15.069 7.601 15.045 / \plot 7.601 15.045 7.690 15.022 / \plot 7.690 15.022 7.781 14.999 / \plot 7.781 14.999 7.874 14.975 / \plot 7.874 14.975 7.967 14.954 / \plot 7.967 14.954 8.062 14.931 / \plot 8.062 14.931 8.160 14.910 / \plot 8.160 14.910 8.259 14.889 / \plot 8.259 14.889 8.361 14.870 / \plot 8.361 14.870 8.462 14.848 / \plot 8.462 14.848 8.568 14.829 / \plot 8.568 14.829 8.674 14.810 / \plot 8.674 14.810 8.784 14.789 / \plot 8.784 14.789 8.867 14.776 / \plot 8.867 14.776 8.951 14.762 / \plot 8.951 14.762 9.038 14.747 / \plot 9.038 14.747 9.129 14.734 / \plot 9.129 14.734 9.220 14.719 / \plot 9.220 14.719 9.313 14.704 / \plot 9.313 14.704 9.409 14.692 / \plot 9.409 14.692 9.506 14.677 / \plot 9.506 14.677 9.605 14.662 / \plot 9.605 14.662 9.707 14.647 / \plot 9.707 14.647 9.811 14.635 / \plot 9.811 14.635 9.914 14.620 / \plot 9.914 14.620 10.022 14.605 / \plot 10.022 14.605 10.132 14.590 / \plot 10.132 14.590 10.243 14.577 / \plot 10.243 14.577 10.355 14.563 / \plot 10.355 14.563 10.467 14.548 / \plot 10.467 14.548 10.581 14.533 / \plot 10.581 14.533 10.696 14.520 / \plot 10.696 14.520 10.812 14.506 / \plot 10.812 14.506 10.926 14.493 / \plot 10.926 14.493 11.043 14.478 / \plot 11.043 14.478 11.159 14.465 / \plot 11.159 14.465 11.273 14.450 / \plot 11.273 14.450 11.390 14.438 / \plot 11.390 14.438 11.504 14.425 / \plot 11.504 14.425 11.616 14.412 / \plot 11.616 14.412 11.728 14.400 / \plot 11.728 14.400 11.839 14.389 / \plot 11.839 14.389 11.946 14.376 / \plot 11.946 14.376 12.052 14.366 / \plot 12.052 14.366 12.158 14.353 / \plot 12.158 14.353 12.260 14.343 / \plot 12.260 14.343 12.361 14.332 / \plot 12.361 14.332 12.459 14.323 / \plot 12.459 14.323 12.554 14.313 / \plot 12.554 14.313 12.647 14.304 / \plot 12.647 14.304 12.736 14.294 / \plot 12.736 14.294 12.823 14.285 / \plot 12.823 14.285 12.907 14.277 / \plot 12.907 14.277 12.990 14.271 / \plot 12.990 14.271 13.070 14.262 / \plot 13.070 14.262 13.147 14.256 / \plot 13.147 14.256 13.221 14.247 / \plot 13.221 14.247 13.293 14.241 / \plot 13.293 14.241 13.363 14.235 / \plot 13.363 14.235 13.464 14.226 / \plot 13.464 14.226 13.559 14.216 / \plot 13.559 14.216 13.652 14.209 / \plot 13.652 14.209 13.739 14.201 / \plot 13.739 14.201 13.824 14.194 / \plot 13.824 14.194 13.904 14.190 / \plot 13.904 14.190 13.981 14.186 / \plot 13.981 14.186 14.055 14.182 / \plot 14.055 14.182 14.125 14.180 / \putrule from 14.125 14.180 to 14.190 14.180 \putrule from 14.190 14.180 to 14.254 14.180 \plot 14.254 14.180 14.313 14.182 / \plot 14.313 14.182 14.368 14.186 / \plot 14.368 14.186 14.421 14.192 / \plot 14.421 14.192 14.470 14.199 / \plot 14.470 14.199 14.514 14.207 / \plot 14.514 14.207 14.556 14.218 / \plot 14.556 14.218 14.594 14.230 / \plot 14.594 14.230 14.630 14.245 / \plot 14.630 14.245 14.662 14.262 / \plot 14.662 14.262 14.692 14.279 / \plot 14.692 14.279 14.719 14.300 / \plot 14.719 14.300 14.745 14.321 / \plot 14.745 14.321 14.766 14.345 / \plot 14.766 14.345 14.787 14.370 / \plot 14.787 14.370 14.806 14.398 / \plot 14.806 14.398 14.823 14.427 / \plot 14.823 14.427 14.840 14.457 / \plot 14.840 14.457 14.855 14.491 / \plot 14.855 14.491 14.870 14.525 / \plot 14.870 14.525 14.884 14.563 / \plot 14.884 14.563 14.897 14.603 / \plot 14.897 14.603 14.912 14.647 / \plot 14.912 14.647 14.927 14.694 / \plot 14.927 14.694 14.939 14.745 / \plot 14.939 14.745 14.954 14.800 / \plot 14.954 14.800 14.967 14.861 / \plot 14.967 14.861 14.982 14.925 / \plot 14.982 14.925 14.997 14.997 / \plot 14.997 14.997 15.011 15.073 / \plot 15.011 15.073 15.026 15.155 / \plot 15.026 15.155 15.043 15.242 / \plot 15.043 15.242 15.058 15.337 / \plot 15.058 15.337 15.075 15.437 / \plot 15.075 15.437 15.092 15.541 / \plot 15.092 15.541 15.109 15.649 / \plot 15.109 15.649 15.128 15.756 / \plot 15.128 15.756 15.145 15.867 / \plot 15.145 15.867 15.160 15.972 / \plot 15.160 15.972 15.176 16.076 / \plot 15.176 16.076 15.189 16.169 / \plot 15.189 16.169 15.202 16.256 / \plot 15.202 16.256 15.215 16.328 / \plot 15.215 16.328 15.223 16.389 / \plot 15.223 16.389 15.229 16.436 / \plot 15.229 16.436 15.234 16.470 / \plot 15.234 16.470 15.238 16.493 / \plot 15.238 16.493 15.240 16.504 / \putrule from 15.240 16.504 to 15.240 16.510 }% \linethickness= 0.500pt \setplotsymbol ({\thinlinefont .}) {\color[rgb]{0,0,0}\plot 15.240 11.430 15.238 11.432 / \plot 15.238 11.432 15.232 11.434 / \plot 15.232 11.434 15.221 11.441 / \plot 15.221 11.441 15.206 11.451 / \plot 15.206 11.451 15.183 11.466 / \plot 15.183 11.466 15.153 11.483 / \plot 15.153 11.483 15.115 11.506 / \plot 15.115 11.506 15.069 11.532 / \plot 15.069 11.532 15.018 11.561 / \plot 15.018 11.561 14.961 11.593 / \plot 14.961 11.593 14.897 11.627 / \plot 14.897 11.627 14.829 11.663 / \plot 14.829 11.663 14.757 11.699 / \plot 14.757 11.699 14.683 11.735 / \plot 14.683 11.735 14.609 11.771 / \plot 14.609 11.771 14.531 11.805 / \plot 14.531 11.805 14.450 11.841 / \plot 14.450 11.841 14.368 11.874 / \plot 14.368 11.874 14.285 11.906 / \plot 14.285 11.906 14.199 11.938 / \plot 14.199 11.938 14.110 11.970 / \plot 14.110 11.970 14.019 12.002 / \plot 14.019 12.002 13.921 12.033 / \plot 13.921 12.033 13.822 12.063 / \plot 13.822 12.063 13.716 12.095 / \plot 13.716 12.095 13.604 12.126 / \plot 13.604 12.126 13.485 12.158 / \plot 13.485 12.158 13.360 12.190 / \plot 13.360 12.190 13.229 12.224 / \plot 13.229 12.224 13.149 12.243 / \plot 13.149 12.243 13.068 12.264 / \plot 13.068 12.264 12.984 12.283 / \plot 12.984 12.283 12.897 12.304 / \plot 12.897 12.304 12.806 12.325 / \plot 12.806 12.325 12.713 12.349 / \plot 12.713 12.349 12.617 12.370 / \plot 12.617 12.370 12.520 12.391 / \plot 12.520 12.391 12.421 12.414 / \plot 12.421 12.414 12.317 12.438 / \plot 12.317 12.438 12.211 12.461 / \plot 12.211 12.461 12.103 12.486 / \plot 12.103 12.486 11.993 12.509 / \plot 11.993 12.509 11.881 12.535 / \plot 11.881 12.535 11.764 12.558 / \plot 11.764 12.558 11.648 12.584 / \plot 11.648 12.584 11.527 12.609 / \plot 11.527 12.609 11.407 12.634 / \plot 11.407 12.634 11.284 12.662 / \plot 11.284 12.662 11.161 12.687 / \plot 11.161 12.687 11.034 12.713 / \plot 11.034 12.713 10.907 12.740 / \plot 10.907 12.740 10.780 12.766 / \plot 10.780 12.766 10.653 12.791 / \plot 10.653 12.791 10.524 12.819 / \plot 10.524 12.819 10.395 12.844 / \plot 10.395 12.844 10.266 12.869 / \plot 10.266 12.869 10.137 12.897 / \plot 10.137 12.897 10.008 12.922 / \plot 10.008 12.922 9.881 12.948 / \plot 9.881 12.948 9.754 12.973 / \plot 9.754 12.973 9.627 12.996 / \plot 9.627 12.996 9.504 13.022 / \plot 9.504 13.022 9.379 13.045 / \plot 9.379 13.045 9.258 13.068 / \plot 9.258 13.068 9.138 13.092 / \plot 9.138 13.092 9.021 13.115 / \plot 9.021 13.115 8.905 13.136 / \plot 8.905 13.136 8.791 13.159 / \plot 8.791 13.159 8.680 13.180 / \plot 8.680 13.180 8.570 13.200 / \plot 8.570 13.200 8.465 13.221 / \plot 8.465 13.221 8.361 13.240 / \plot 8.361 13.240 8.259 13.259 / \plot 8.259 13.259 8.160 13.276 / \plot 8.160 13.276 8.064 13.295 / \plot 8.064 13.295 7.969 13.312 / \plot 7.969 13.312 7.878 13.329 / \plot 7.878 13.329 7.791 13.343 / \plot 7.791 13.343 7.705 13.358 / \plot 7.705 13.358 7.622 13.373 / \plot 7.622 13.373 7.542 13.388 / \plot 7.542 13.388 7.415 13.409 / \plot 7.415 13.409 7.294 13.430 / \plot 7.294 13.430 7.178 13.449 / \plot 7.178 13.449 7.065 13.468 / \plot 7.065 13.468 6.960 13.485 / \plot 6.960 13.485 6.856 13.500 / \plot 6.856 13.500 6.759 13.515 / \plot 6.759 13.515 6.663 13.528 / \plot 6.663 13.528 6.574 13.538 / \plot 6.574 13.538 6.488 13.547 / \plot 6.488 13.547 6.405 13.555 / \plot 6.405 13.555 6.327 13.559 / \plot 6.327 13.559 6.253 13.564 / \plot 6.253 13.564 6.183 13.566 / \putrule from 6.183 13.566 to 6.115 13.566 \plot 6.115 13.566 6.054 13.564 / \plot 6.054 13.564 5.994 13.561 / \plot 5.994 13.561 5.941 13.555 / \plot 5.941 13.555 5.891 13.547 / \plot 5.891 13.547 5.844 13.538 / \plot 5.844 13.538 5.802 13.526 / \plot 5.802 13.526 5.762 13.513 / \plot 5.762 13.513 5.726 13.498 / \plot 5.726 13.498 5.692 13.481 / \plot 5.692 13.481 5.660 13.462 / \plot 5.660 13.462 5.632 13.441 / \plot 5.632 13.441 5.607 13.420 / \plot 5.607 13.420 5.584 13.394 / \plot 5.584 13.394 5.563 13.369 / \plot 5.563 13.369 5.541 13.341 / \plot 5.541 13.341 5.522 13.314 / \plot 5.522 13.314 5.503 13.282 / \plot 5.503 13.282 5.482 13.244 / \plot 5.482 13.244 5.463 13.204 / \plot 5.463 13.204 5.444 13.159 / \plot 5.444 13.159 5.425 13.113 / \plot 5.425 13.113 5.406 13.060 / \plot 5.406 13.060 5.389 13.003 / \plot 5.389 13.003 5.370 12.939 / \plot 5.370 12.939 5.353 12.871 / \plot 5.353 12.871 5.334 12.797 / \plot 5.334 12.797 5.315 12.717 / \plot 5.315 12.717 5.296 12.630 / \plot 5.296 12.630 5.277 12.537 / \plot 5.277 12.537 5.256 12.438 / \plot 5.256 12.438 5.237 12.334 / \plot 5.237 12.334 5.215 12.228 / \plot 5.215 12.228 5.196 12.118 / \plot 5.196 12.118 5.177 12.010 / \plot 5.177 12.010 5.158 11.904 / \plot 5.158 11.904 5.141 11.803 / \plot 5.141 11.803 5.127 11.712 / \plot 5.127 11.712 5.112 11.631 / \plot 5.112 11.631 5.101 11.565 / \plot 5.101 11.565 5.093 11.513 / \plot 5.093 11.513 5.086 11.474 / \plot 5.086 11.474 5.082 11.449 / \plot 5.082 11.449 5.080 11.436 / \putrule from 5.080 11.436 to 5.080 11.430 }% \put{\SetFigFont{6}{7.2}{\rmdefault}{\mddefault}{\updefault}{\color[rgb]{0,0,0}$\tilde{S}_-$}% } [lB] at 3.810 16.510 \put{\SetFigFont{6}{7.2}{\rmdefault}{\mddefault}{\updefault}{\color[rgb]{0,0,0}$\bar{N}_-$}% } [lB] at 3.810 11.430 \put{\SetFigFont{6}{7.2}{\rmdefault}{\mddefault}{\updefault}{\color[rgb]{0,0,0}$\tilde{N}_-$}% } [lB] at 3.810 21.590 \put{\SetFigFont{6}{7.2}{\rmdefault}{\mddefault}{\updefault}{\color[rgb]{0,0,0}$\tilde{S}_+$}% } [lB] at 15.558 11.430 \put{\SetFigFont{6}{7.2}{\rmdefault}{\mddefault}{\updefault}{\color[rgb]{0,0,0}$\tilde{N}_+$}% } [lB] at 15.558 16.510 \put{\SetFigFont{6}{7.2}{\rmdefault}{\mddefault}{\updefault}{\color[rgb]{0,0,0}$\bar{N}_+$}% } [lB] at 15.558 6.350 \put{\SetFigFont{6}{7.2}{\rmdefault}{\mddefault}{\updefault}{\color[rgb]{0,0,0}$\Gamma_l$}% } [lB] at 5.715 12.700 \put{\SetFigFont{6}{7.2}{\rmdefault}{\mddefault}{\updefault}{\color[rgb]{0,0,0}$\Gamma_r$}% } [lB] at 14.287 14.605 \put{\SetFigFont{6}{7.2}{\rmdefault}{\mddefault}{\updefault}{\color[rgb]{0,0,0}${\mathcal S}^\Theta(E,\lambda)$}% } [lB] at 9.525 13.494 \put{\SetFigFont{6}{7.2}{\rmdefault}{\mddefault}{\updefault}{\color[rgb]{0,0,0}$\mathcal{R}$}% } [lB] at 11.748 15.558 \put{\SetFigFont{6}{7.2}{\rmdefault}{\mddefault}{\updefault}{\color[rgb]{0,0,0}$\mathcal{W}^-(E',\lambda)$}% } [lB] at 7.620 15.081 \linethickness=0pt \putrectangle corners at 3.778 21.901 and 15.589 6.215 \endpicture} \end{center} \caption{Trapped region $\mathcal{R}$ and the corridor with zero winding number.} \end{figure} Consider the region $\mathcal{R}$ in $\bar{\mathcal{C}}$ whose boundary consists of the following three curves: (i) From $S_-$ to $S_+$ along $\Theta_{E,\lambda}$. (ii) From $S_+$ to $N_+$ along the right boundary, and (iii) from $N_+$ to $S_-$ along a horizontal line segment. We claim that $\mathcal{R}$ is a trapped region for \refeq{dynsysTh} at parameter values $(E',\lambda)$ provided $E'<E$. Since (ii) is always an orbit, and (iii) is entirely contained in ${\mathcal N}$, this only needs to be checked for (i). We have $$ g_{E',\lambda}(\theta,\Theta_{E,\lambda}) = g_{E,\lambda}(\theta,\Theta_{E,\lambda}) + 2a(E'-E)\sin^2\theta \sin\Theta_{E,\lambda} \geq \dot{\Theta}_{E,\lambda} $$ which shows that the new flow crosses the old solution from left to right. Thus $\mathcal{R}$ is trapped and since $\mathcal{W}^-$ starts in $\mathcal{R}$, it must terminate in $\widetilde{N}_+ = (\pi,0)$, so that $w(\mathcal{W}^-) = 0$. \end{proof} Setting $E_0 =1$ and $\lambda_0 = -1+a$, let ${\mathcal S}_0 := {\mathcal S}(E_0,\lambda_0)$ denote the explicit solution found in \ref{explicit}. For all $E_1\in (0,1)$, the above two propositions, together with the following immediate corollary of Proposition~\ref{prop:ssexists}, establish the existence of a saddles connector ${\mathcal S}_1 := {\mathcal S}(E_1,\lambda_1)$ for \refeq{dynsysTh}, for some $\lambda_1\in (\lambda_l,\lambda_0)$: \begin{cor}\label{cor:thconex} Let $E\in[0,1]$ be fixed. Suppose that there exists $\lambda_1< \lambda_2<0$ such that the flow \refeq{dynsysTh} has corridors ${\mathcal K}_1(E,\lambda_1)$ and ${\mathcal K}_1(E,\lambda_2)$ with $w({\mathcal K}_1(E,\lambda_2))=0$ and $w({\mathcal K}_1(E,\lambda_1))\geq 1$. Then there is a $\lambda\in(\lambda_1,\lambda_2)$ such that \refeq{dynsysTh} has a saddles connector ${\mathcal S}^\Theta(E,\lambda)$ going from $(0,0)$ to $(\pi,-\pi)$. \end{cor} \begin{proof} Proposition~\ref{prop:ssexists} applies, with $-\lambda$ playing the role of the parameter $\mu$. \end{proof} Proceeding iteratively, suppose that given $E_n\in[0,1]$ a saddles connector ${\mathcal S}^\Theta_n={\mathcal S}^\Theta(E_n,\lambda_n)$ has been found for \refeq{dynsysTh}, for some $\lambda_n <\lambda_0$. In the next subsection we shall see how the newly-found $\lambda_n$ can be used to prove the existence of a saddles connector for the $\Omega$ flow \refeq{dynsysOm}, namely ${\mathcal S}^\Omega_n:= {\mathcal S}^\Omega_n(E_{n+1},\lambda_n)$ for some $E_{n+1} \in (0,1)$. Coming back to the $\Theta$ flow then, a new saddles connector ${\mathcal S}_{n+1}^\Theta$ needs to be found with the updated energy $E_{n+1}$, {\em given that a saddles connector ${\mathcal S}^\Theta_n(E_n,\lambda_n)$ already exists.} Since $E_{n+1}$ can be on either side of $E_n$, in addition to Prop.~\ref{prop:thlowercor} we also need the following: \begin{thm}\label{thm:La} Given any $a\in(0,\frac{1}{2})$ and $E \in [0,1]$, there exists a unique $$ \lambda = \Lambda(E) \in [-1-a,-1+a] $$ such that \refeq{dynsysTh} has a saddles connector ${\mathcal S}(E,\lambda)$ going from $(0,0)$ to $(\pi,-\pi)$. Moreover, $\Lambda$ is an increasing $C^1$ function, and $\frac{\partial \Lambda}{\partial E} < a$. \end{thm} \begin{proof} If $E=1$ then $\lambda = -1+a$ works and ${\mathcal S} = {\mathcal S}_0 = (\theta,-\theta)$. For $E<1$ existence is guaranteed by Prop.~\ref{prop:thlowercor}, Prop.~\ref{prop:lacorlow}, and Corollary~\ref{cor:thconex}. To prove uniqueness, suppose that for a given $E$, there are two saddles connectors ${\mathcal S}(E,\lambda)$ and ${\mathcal S}'(E,\lambda')$, with $\lambda'>\lambda$. Let $\Theta_{E,\lambda}$ and $\Theta_{E,\lambda'}$ denote the corresponding $\Theta$-components of ${\mathcal S}$ and ${\mathcal S}'$, respectively. For $\theta\in(0,\pi)$, $$ g_{E,\lambda'}(\theta,\Theta_{E,\lambda}) = g_{E,\lambda}(\theta,\Theta_{E,\lambda}) + 2(\lambda'-\lambda)\sin\theta > \dot{\Theta}_{E,\lambda}. $$ Thus orbits of the $(E,\lambda')$ flow can only cross ${\mathcal S}$ from below to above. On the other hand, since ${\mathcal S}'$ is a saddles connector, near $S_-$ it coincides with $\mathcal{W}^-(E,\lambda')$, and near $S_+$ it coincides with $\mathcal{W}^+$. Thus from the linearizartion of the flow at $S_\pm$, $$ \left.\frac{d\Theta_{E,\lambda'}}{d\theta}\right|_{S_\pm} = \lambda'-a>\lambda-a. $$ Therefore ${\mathcal S}'$ must be above ${\mathcal S}$ near $S_-$ and below it near $S_+$, so ${\mathcal S}'$ would have to cross ${\mathcal S}$ from above to below, which is a contradiction, unless they coincide. Given $E$ then, let $\Lambda(E)$ denote the unique value of $\lambda$ for which a saddles connector ${\mathcal S}(E,\lambda)$ exists. The fact that $\Lambda$ is continuously differentiable (in fact analytic), and the bound on the derivative, have already been shown in \cite{WinklmeierPHD} and \cite{WINKLMEIERa} using analytic perturbation theory. Here we give a simple proof of monotonicity of $\Lambda$ which also establishes the bound on the derivative: Given $E\in[0,1]$ let $\lambda=\Lambda(E)$ and let $(\theta(\tau),\Theta_E(\tau))$ denote the unique (modulo translations in $\tau$) saddles connector for \refeq{dynsysTh} whose existence we have established. Let $u: = \frac{\partial \Theta_E}{\partial E}$. By differentiating the $\Theta$ equation in \refeq{dynsysTh} with respect to $E$ we obtain an equation for $u$: \begin{equation}\label{eq:uu} \frac{du}{d\tau} = P(\tau) u + Q(\tau),\qquad \left\{\begin{array}{rcl} P & := & 2a\sin\theta\cos\theta \sin\Theta_E + (2aE\sin^2\theta - 1)\cos\Theta_E \\ Q & := & 2a\sin^2\theta\sin\Theta_E + 2 \frac{d\Lambda}{dE} \sin\theta.\end{array}\right. \end{equation} Let $$ U(\tau_1^{},\tau_2^{}) := e^{-\int_{\tau_1^{}}^{\tau_2^{}} P(\tau) d\tau}. $$ Thus we have $$ U(\tau_2^{},\tau_1^{}) = \frac{1}{U(\tau_1^{},\tau_2^{})},\qquad U(\tau_1^{},\tau_2^{})U(\tau_2^{},\tau_3^{}) = U(\tau_1^{},\tau_3^{}). $$ Solving the first-order linear ODE \refeq{eq:uu} for $u$ we obtain \begin{equation}\label{UQ} u(\tau) = U(\tau,\tau_1^{}) u(\tau_1^{}) + \int_{\tau_1^{}}^\tau U(\tau,\tau') Q(\tau') d\tau'. \end{equation} Note that $P$ and $Q$ are bounded functions of $\tau$ and $$ \lim_{\tau\to -\infty} P(\tau) = -1,\qquad \lim_{\tau\to\infty} P(\tau) = 1. $$ Therefore, for any fixed $\tau$, $$ U(\tau,\tau_1^{}) \to 0\mbox{ as }\tau_1^{}\to -\infty,\qquad U(\tau,\tau_1^{}) \to 0\mbox{ as } \tau_1^{}\to \infty. $$ Moreover $u(\pm\infty) = 0$. For any finite $\tau$ from \refeq{UQ} we thus obtain two equivalent expressions for $u(\tau)$. For example, setting $\tau=0$, $$ u(0) = \int_{-\infty}^0 U(0,\tau') Q(\tau') d\tau' =- \int_0^\infty U(0,\tau') d\tau'. $$ Thus in particular $$ 0 = \int_{-\infty}^\infty U(0,\tau') Q(\tau') d\tau' = 2a\int_{\mathbb R} U(0,\tau')\sin^2\theta(\tau')\sin\Theta_E(\tau') d\tau' + 2 \frac{d\Lambda}{d E} \int_{\mathbb R} U(0,\tau') \sin\theta(\tau') d\tau' $$ Therefore $$ \frac{d\Lambda}{dE} =a \frac{\int_{\mathbb R} U(0,\tau')\sin^2\theta(\tau')\left(-\sin\Theta_E(\tau')\right) d\tau'}{\int_{\mathbb R} U(0,\tau') \sin\theta(\tau') d\tau'} \geq 0. $$ We have already shown that $\Theta_E(\tau)\in(-\pi,0)$ and $\theta(\tau)\in(0,\pi)$ for all $\tau$. Thus the numerator in the above fraction is strictly less than the denominator, hence \begin{equation}\label{est:dlade} 0\leq \frac{\partial\Lambda}{\partial E} < a \end{equation} \end{proof} \subsubsection{Existence of saddles connectors for the $\Omega$ equation} We now show that the flow \refeq{dynsysOm} also satisfies all the hypotheses we had made about flows on a cylinder. Once again, for simplicity we are only going to consider the case $\kappa =\frac{1}{2}$. The situation is somewhat more complicated than what we have done in the above for the $\Theta$ equation, due to the presence of an extra parameter, namely $\gamma$, which breaks the symmetry that was present for the $\Theta$ flow, as well as the fact that the equilibria of the $\Omega$ flow are degenerate (non-hyperbolic). Let us make the identifications $x = \xi$ and $y=\Omega$. Thus $x_-^{} = -\pi/2$, $x_+^{} = \pi/2$, and we now have $$ f(\xi) = \cos^2 \xi, \qquad g_{E,\lambda}(\xi,\Omega) = 2a\sin\xi\cos\Omega+2\lambda\cos\xi\sin\Omega + 2\gamma\sin\xi\cos\xi+\cos^2\xi - 2aE $$ Therefore, for $E\in[0,1)$, $$ s_-^{} = -\pi+\cos^{-1}(E),\quad n_-^{} = \pi - \cos^{-1}(E),\qquad s_+^{} = -\cos^{-1}(E),\quad n_+^{} = \cos^{-1}(E), $$ where by $\cos^{-1}$ we mean the principal branch of the arccosine, $0\leq \cos^{-1}x\leq \pi$, and $$ S_\pm = (\pm{\textstyle\frac{\pi}{2}},s_\pm),\qquad N_\pm = (\pm{\textstyle\frac{\pi}{2}},n_\pm) $$ as before. We note that this time, all the equilibria are non-hyperbolic, since $f'(\pm\pi/2) = 0$, and that for $\gamma= 0$, there is a discrete symmetry: $f(-\xi) = f(\xi)$ and $g_{|_{\gamma=0}}(-\xi,\pi-\Omega) = g_{|_{\gamma=0}}(\xi,\Omega)$, which is broken when $\gamma$ is turned on. Also in the case $E=1$ there is a further degeneracy: the two equilibria on each side coalesce into one singular point with both eigenvalues equal to zero. For these type of singular points center manifolds can be non-unique, so that the distinguished orbits $\mathcal{W}^\pm$ and the index theory we have developed for the corridor they form, are not directly relevent to this case. We now check the hypotheses about the topology of the nullclines. \subsubsection{Topology of the null-clines} Let $T := \tan(\Omega/2)$. Then $$ g_{E,\lambda}(\xi,\Omega) = \frac{2q(T)}{1+T^2}, $$ where \begin{eqnarray*} q(T) & := & \left(\gamma\sin\xi\cos\xi +{\textstyle\frac{1}{2}}\cos^2\xi - aE -a\sin\xi\right)T^2 + 2\lambda\cos\xi T\\ &&\mbox{} + \left(\gamma\sin\xi\cos\xi +{\textstyle\frac{1}{2}}\cos^2\xi - aE +a\sin\xi\right). \end{eqnarray*} Thus $q$ is a quadratic polynomial in $T$ with coefficients that are functions of $\xi$. The discriminant of $q$ is $$ \Delta_q(\xi) := \lambda^2\cos^2\xi - (\gamma\sin\xi\cos\xi + {\textstyle\frac{1}{2}}\cos^2\xi - aE)^2 + 4a^2\sin^2\xi. $$ Let $\tau := \tan\xi$. Thus $-\infty<\tau<\infty$, and $$ \Delta_q(\tau) = \frac{p(\tau)}{(1+\tau^2)^2} $$ with $$ p(\tau) := a^2(1-E^2) \tau^4 + 2\gamma a E \tau^3 + \left( \lambda^2 -\gamma^2 + a^2 + 2a({\textstyle\frac{1}{2}} - aE)\right) \tau^2 - 2 \gamma({\textstyle\frac{1}{2}} - aE) \tau + \lambda^2 - ({\textstyle\frac{1}{2}} - aE)^2. $$ Since $ |E|<1$, $p$ is an irreducible quartic in $\tau$. Let us write it as $p(\tau) = \sum_{j=0}^4 c_j \tau^j$. Suppose that $E\in(0,1)$ and \begin{equation}\label{paramrange} a \in (0,{\textstyle\frac{1}{2}}),\qquad \gamma \in (- \sqrt{2a(1-2a)},0),\qquad \lambda \in [-1-a,-1+a] \end{equation} It then follows that $$ c_4>0,\qquad c_3<0,\qquad c_2>0,\qquad c_1>0,\qquad c_0>0. $$ By Descartes' Rule of Signs, then, $p(\tau)$ has either two or no real positive roots, and either two or no real negative roots, counting multiplicity. For a more accurate root count, one needs to use the discriminant of the quartic. Since the discriminant theory for general quartics is somewhat complicated, here we opt for a simpler analysis by estimating $p$ from above and below with two {\em reducible} quartics. To this end, first we note that $ p(\tau) = Q(\tau) - (q_1(\tau))^2 $ with $$ Q(\tau) = \lambda^2 + (\lambda^2+a^2) \tau^2 + a^2 \tau^4,\qquad q_1(\tau) = {\textstyle\frac{1}{2}} - aE + \gamma \tau - aE \tau^2. $$ Thus on the one hand, by completing the square, $$ Q(\tau) = \left(a\tau^2+\frac{\lambda^2+a^2}{2a}\right)^2 - \frac{(\lambda^2 - a^2)^2}{4a^2} \leq (q_2^+(\tau))^2,\qquad q_2^+(\tau) := a\tau^2 + \frac{\lambda^2+a^2}{2a} $$ and on the other hand, $$ Q(\tau) \geq \lambda^2 + 2 |\lambda| a \tau^2 + a^2 \tau^4 = (q_2^-(\tau))^2,\qquad q_2^-(\tau) = a\tau^2 + |\lambda| $$ Thus we have upper and lower bounds for $p$ in terms of factorizable quartics $Q^\pm(\tau)$: $$ Q^-(\tau) := \left(q_2^-(\tau)\right)^2 - \left(q_1(\tau)\right)^2 \leq p(\tau) \leq \left(q_2^+(\tau)\right)^2 - \left(q_1(\tau)\right)^2 =: Q^+(\tau) $$ \begin{figure}[ht] \begin{center} \includegraphics[scale =0.3]{KTZ_zGKN_I_quartics1} \includegraphics[scale=0.3]{KTZ_zGKN_I_quartics2} \end{center} \caption{The bounding quartics: $Q^+$ and $p$ (left), $p$ and $Q^-$ (right) for parameter values $a=0.1$, $\gamma=-0.2$, $\lambda = -0.9$, $E = 0.978$.} \end{figure} Consider first the upper quartic $Q^+$. We have $Q^+ = (q_2^+ + q_1)(q_2^+ - q_1) = q^+ q^-$ where $q^\pm$ are two quadratic polynomials. It is clear that if $q^+$ has any real roots, they will be positive, and if $q^-$ has any roots, they will be negative (recall that we are assuming $\gamma<0$). The discriminants of $q^\pm$ are computed to be $$ \Delta_\pm := \gamma^2 - 2 (1 \mp E)\left( \lambda^2 + a^2 \pm a(1 -2aE)\right). $$ Similarly, $Q^- = (q_2^-+q_1)(q_2^--q_1) = \tilde{q}^+\tilde{q}^-$ where $\tilde{q}^\pm$ are two quadratics, and once again, any real roots of $\tilde{q}^+$ must be positive and any real root of $\tilde{q}^-$, negative. The discriminants of $\tilde{q}^\pm$ are $$ \tilde{\Delta}_\pm = \gamma^2 - 2(1\mp E) (-2a\lambda \pm a(1 - 2aE) )\geq \Delta_\pm $$ It thus follows that there are two subsets of the $(a,\gamma,\lambda)$ parameter space \refeq{paramrange} that are of interest: (R1) where both $\tilde{\Delta}_+$ and $\tilde{\Delta}_-$ are negative; and (R2) where $\Delta_+>0$ and $\tilde{\Delta}_-<0$. For parameter values in the region (R1) the quartic $Q^-$ will have no real zeros, and will be always positive, while for those in (R2) the quartic $Q^+$ will have exactly two positive roots and no negative root. Let us fix $a,\gamma,\lambda$ as in \refeq{paramrange}. We find that the range (R1) corresponds to $0\leq E < E_l$, where $$ E_l(\lambda) := \frac{1}{2a}\left[ -\lambda+a+{\textstyle\frac{1}{2}} - \sqrt{(\lambda+a-{\textstyle\frac{1}{2}})^2+\gamma^2}\right], $$ while range (R2) corresponds to $E_h< E\leq 1$, with $$ E_h(\lambda) = \frac{1}{4a^2}\left[\lambda^2+3a^2+a - \sqrt{(\lambda^2 - a^2 + a)^2 + 4 a^2 \gamma^2}\right] $$ Note that $0< E_l<E_h<1$ for all values of $a,\gamma,\lambda$ as in \refeq{paramrange}. For $E\in(E_h,1]$, therefore, since $Q^+$ has two positive roots, the quartic $p(\tau)$ must also have at least two roots, one of which will definitely be positive. Thus by the Rule of Signs, $p(\tau)$ has exactly two positive roots. We call them $\tau_1^{}$ and $\tau_2^{}$, and $p(\tau)<0$ for $\tau_1^{}<\tau<\tau_2^{}$. It follows that the quadratic $q(T)$ will have two roots for $\tau\notin[\tau_1^{},\tau_2^{}]$, double roots at $\tau_1^{}$ and at $\tau_2^{}$, and no real roots for $\tau\in(\tau_1^{},\tau_2^{})$. For $E\in[0,E_l)$ since $Q^-$ has no real roots, $p$ cannot have any either. Thus $p$ is always positive and $q(T)=0$ will have two roots for all $\tau\in {\mathbb R}$. Combining these two, one concludes that a critical value for the energy $E=E_c(a,\gamma,\lambda)$ exists, $E_c\in[E_l,E_h]$, such that assumption ({\bf A}) is satisfied, with the role of parameter $\mu$ played by $E$. \begin{figure}[ht] \begin{center} \includegraphics[scale=0.3]{KTZ_zGKN_I_omeganull1} \includegraphics[scale=0.3]{KTZ_zGKN_I_omeganull2} \end{center} \caption{$\Omega$-nullclines for parameter values $E=0.8$ (left) and $E=0.93$ (right), with $a=0.1$, $\gamma = -0.4$, and $\lambda = -0.9$.} \end{figure} \subsubsection{Existence of corridors with unequal winding number} Throughout this section, $a$ will be a fixed number in $(0,\frac{1}{2})$ and $\gamma$ a fixed number in $(- \sqrt{2a(1-2a)},0)$. The following two propositions help us get started: \begin{prop}\label{prop:om1} Given $\lambda \leq -1 +a$ there exists $\bar{E}\in(E_h(\lambda),1)$ such that for all $E\in[\bar{E},1)$ the corridor ${\mathcal K}_1(E,\lambda)$ of the flow \refeq{dynsysOm} has winding number greater than or equal to one. \end{prop} \begin{proof} We compute the slope of solution orbits that cross the following line in $\bar{\mathcal{C}}$ $$L := \{(\xi,\Omega)\in \bar{\mathcal{C}}\ |\ -{\textstyle\frac{\pi}{2}}\leq\xi\leq{\textstyle\frac{\pi}{2}},\quad \Omega = {\textstyle\frac{\pi}{2}} - \cos^{-1}E - \xi\}$$ and compare it to the slope of $L$. Note that $L$ passes through $\widetilde{N}_-$ and $\widetilde{S}_+$. We have \begin{equation}\label{slopes} \frac{g_{E,\lambda}(\xi,\Omega)}{f(\xi)} - (-1) = 2\left( 1 - (a-\lambda)E + [ \sqrt{1-E^2} (a-\lambda) + \gamma]\tan\xi \right) \end{equation} Consider first the case $\lambda = -1+a$. Let $$ \xi_0(E) := \tan^{-1} \frac{1 - E}{-\gamma-\sqrt{1-E^2}} \searrow 0 \mbox{ as } E \nearrow 1. $$ For $\xi\geq \xi_0(E)$, the slope of any orbit of the flow crossing $L$ is less than (i.e. more negative than) the slope of $L$. Hence on the portion of $L$ where $\xi\geq \xi_0$ orbits can only cross $L$ from above to below. On the other hand, suppose $\lambda<-1+a$. Let $$ E_m := \max\left\{\frac{1}{a-\lambda},\sqrt{1-\frac{\gamma^2}{(a-\lambda)^2}},E_h(\lambda)\right\}\in(E_h,1) $$ For $E\in (E_m,1)$ we have $\xi_0(E)\in(-\frac{\pi}{2},0)$. Let $$ \eta_0 := \left\{\begin{array}{lr} \cos^{-1}E & \lambda<-1+a \\ \cos^{-1}E +\xi_0(E) & \lambda =-1+a \end{array}\right. $$ We note that $\eta_0 \to 0$ as $E\to 1$. Let us consider the horizontal line $$ L' := \{(\xi,\Omega)\in\bar{\mathcal{C}}\ |\ -{\textstyle\frac{\pi}{2}}\leq \xi\leq {\textstyle\frac{\pi}{2}},\quad \Omega = {\textstyle\frac{\pi}{2}} - \eta_0\} $$ We compute the slope of orbits crossing $L'$: $$ h(\xi) := g_{E,\lambda}(\xi,{\textstyle\frac{\pi}{2}}-\eta_0) = 2a\sin\xi \sin\eta_0 + 2\lambda\cos\xi\cos\eta_0 + 2\gamma \sin\xi\cos\xi + \cos^2\xi -2aE. $$ Clearly \begin{eqnarray*} h(\xi) & \leq & 2a(\sin\eta_0 - E)+|\gamma| +2\lambda\cos\xi\cos\eta_0+ \cos^2\xi \\ & \leq & 2a(\sin\eta_0 - E)+|\gamma| + 2\lambda\cos\eta_0 +1 \\ & \to & -2a + |\gamma| + 2\lambda + 1\quad\mbox{as } E\to 1\\ & \leq & |\gamma| - 1 < 0 \end{eqnarray*} Thus there exists $E$ sufficiently close to 1 such that $h(\xi)<0$ for $\xi\in(-{\textstyle\frac{\pi}{2}},{\textstyle\frac{\pi}{2}})$. \begin{figure}[ht] \begin{center} \font\thinlinefont=cmr5 \begingroup\makeatletter\ifx\SetFigFont\undefined% \gdef\SetFigFont#1#2#3#4#5{% \reset@font\fontsize{#1}{#2pt}% \fontfamily{#3}\fontseries{#4}\fontshape{#5}% \selectfont}% \fi\endgroup% \mbox{\beginpicture \setcoordinatesystem units <0.50000cm,0.50000cm> \unitlength=0.50000cm \linethickness=1pt \setplotsymbol ({\makebox(0,0)[l]{\tencirc\symbol{'160}}}) \setshadesymbol ({\thinlinefont .}) \setlinear \linethickness= 0.500pt \setplotsymbol ({\thinlinefont .}) {\color[rgb]{0,0,0}\putrule from 5.080 11.430 to 15.240 11.430 }% \linethickness=1pt \setplotsymbol ({\makebox(0,0)[l]{\tencirc\symbol{'160}}}) {\color[rgb]{0,0,0}\putrule from 5.080 21.590 to 15.240 21.590 }% \linethickness= 0.500pt \setplotsymbol ({\thinlinefont .}) {\color[rgb]{0,0,0}\putrule from 5.080 17.462 to 5.080 16.034 \plot 5.017 16.288 5.080 16.034 5.143 16.288 / }% \linethickness= 0.500pt \setplotsymbol ({\thinlinefont .}) {\color[rgb]{0,0,0}\putrule from 15.240 12.383 to 15.240 10.795 \plot 15.176 11.049 15.240 10.795 15.304 11.049 / }% \linethickness= 0.500pt \setplotsymbol ({\thinlinefont .}) {\color[rgb]{0,0,0}\putrule from 5.080 16.510 to 15.240 16.510 }% \linethickness=1pt \setplotsymbol ({\makebox(0,0)[l]{\tencirc\symbol{'160}}}) {\color[rgb]{0,0,0}\putrule from 15.240 21.590 to 15.240 6.315 \putrule from 15.240 6.350 to 5.045 6.350 \putrule from 5.080 6.350 to 5.080 21.590 }% \linethickness= 0.500pt \setplotsymbol ({\thinlinefont .}) {\color[rgb]{0,0,0}\plot 15.240 15.875 5.080 20.955 / }% \linethickness= 0.500pt \setplotsymbol ({\thinlinefont .}) {\color[rgb]{0,0,0}\putrule from 15.240 18.415 to 5.080 18.415 }% \linethickness= 0.500pt \setplotsymbol ({\thinlinefont .}) {\color[rgb]{0,0,0}\putrule from 5.080 18.415 to 10.160 18.415 \plot 10.160 18.415 15.240 15.875 / }% \linethickness= 0.500pt \setplotsymbol ({\thinlinefont .}) {\color[rgb]{0,0,0}\plot 6.032 18.733 6.668 18.098 / \plot 6.443 18.232 6.668 18.098 6.533 18.322 / }% \linethickness= 0.500pt \setplotsymbol ({\thinlinefont .}) {\color[rgb]{0,0,0}\plot 8.255 18.733 8.890 18.098 / \plot 8.665 18.232 8.890 18.098 8.755 18.322 / }% \linethickness= 0.500pt \setplotsymbol ({\thinlinefont .}) {\color[rgb]{0,0,0}\plot 11.430 17.939 11.906 17.304 / \plot 11.703 17.469 11.906 17.304 11.805 17.545 / }% \linethickness= 0.500pt \setplotsymbol ({\thinlinefont .}) {\color[rgb]{0,0,0}\plot 13.335 16.986 13.652 16.351 / \plot 13.482 16.550 13.652 16.351 13.596 16.607 / }% \linethickness= 0.500pt \setplotsymbol ({\thinlinefont .}) {\color[rgb]{0,0,0}\plot 15.240 15.875 15.234 15.879 / \plot 15.234 15.879 15.221 15.890 / \plot 15.221 15.890 15.200 15.909 / \plot 15.200 15.909 15.168 15.934 / \plot 15.168 15.934 15.128 15.968 / \plot 15.128 15.968 15.079 16.010 / \plot 15.079 16.010 15.028 16.055 / \plot 15.028 16.055 14.973 16.101 / \plot 14.973 16.101 14.920 16.148 / \plot 14.920 16.148 14.872 16.192 / \plot 14.872 16.192 14.825 16.235 / \plot 14.825 16.235 14.783 16.273 / \plot 14.783 16.273 14.747 16.311 / \plot 14.747 16.311 14.713 16.345 / \plot 14.713 16.345 14.685 16.375 / \plot 14.685 16.375 14.660 16.404 / \plot 14.660 16.404 14.639 16.432 / \plot 14.639 16.432 14.620 16.457 / \plot 14.620 16.457 14.605 16.482 / \plot 14.605 16.482 14.590 16.512 / \plot 14.590 16.512 14.577 16.540 / \plot 14.577 16.540 14.567 16.567 / \plot 14.567 16.567 14.561 16.597 / \plot 14.561 16.597 14.554 16.624 / \plot 14.554 16.624 14.552 16.654 / \putrule from 14.552 16.654 to 14.552 16.681 \plot 14.552 16.681 14.554 16.709 / \plot 14.554 16.709 14.558 16.736 / \plot 14.558 16.736 14.565 16.764 / \plot 14.565 16.764 14.575 16.789 / \plot 14.575 16.789 14.586 16.815 / \plot 14.586 16.815 14.599 16.838 / \plot 14.599 16.838 14.613 16.859 / \plot 14.613 16.859 14.628 16.880 / \plot 14.628 16.880 14.647 16.899 / \plot 14.647 16.899 14.664 16.916 / \plot 14.664 16.916 14.685 16.933 / \plot 14.685 16.933 14.709 16.950 / \plot 14.709 16.950 14.734 16.967 / \plot 14.734 16.967 14.764 16.984 / \plot 14.764 16.984 14.796 16.999 / \plot 14.796 16.999 14.834 17.016 / \plot 14.834 17.016 14.876 17.033 / \plot 14.876 17.033 14.925 17.050 / \plot 14.925 17.050 14.978 17.067 / \plot 14.978 17.067 15.033 17.084 / \plot 15.033 17.084 15.088 17.101 / \plot 15.088 17.101 15.138 17.115 / \plot 15.138 17.115 15.181 17.128 / \plot 15.181 17.128 15.212 17.137 / \plot 15.212 17.137 15.232 17.143 / \plot 15.232 17.143 15.238 17.145 / \putrule from 15.238 17.145 to 15.240 17.145 }% \linethickness= 0.500pt \setplotsymbol ({\thinlinefont .}) {\color[rgb]{0,0,0}\plot 5.080 12.065 5.082 12.067 / \plot 5.082 12.067 5.084 12.076 / \plot 5.084 12.076 5.091 12.088 / \plot 5.091 12.088 5.101 12.107 / \plot 5.101 12.107 5.114 12.137 / \plot 5.114 12.137 5.133 12.175 / \plot 5.133 12.175 5.156 12.224 / \plot 5.156 12.224 5.186 12.283 / \plot 5.186 12.283 5.220 12.353 / \plot 5.220 12.353 5.258 12.431 / \plot 5.258 12.431 5.302 12.520 / \plot 5.302 12.520 5.351 12.617 / \plot 5.351 12.617 5.402 12.721 / \plot 5.402 12.721 5.457 12.831 / \plot 5.457 12.831 5.514 12.946 / \plot 5.514 12.946 5.573 13.064 / \plot 5.573 13.064 5.635 13.183 / \plot 5.635 13.183 5.698 13.303 / \plot 5.698 13.303 5.759 13.424 / \plot 5.759 13.424 5.823 13.542 / \plot 5.823 13.542 5.884 13.659 / \plot 5.884 13.659 5.946 13.771 / \plot 5.946 13.771 6.005 13.881 / \plot 6.005 13.881 6.064 13.989 / \plot 6.064 13.989 6.121 14.091 / \plot 6.121 14.091 6.179 14.190 / \plot 6.179 14.190 6.234 14.285 / \plot 6.234 14.285 6.289 14.376 / \plot 6.289 14.376 6.342 14.463 / \plot 6.342 14.463 6.394 14.548 / \plot 6.394 14.548 6.447 14.628 / \plot 6.447 14.628 6.498 14.704 / \plot 6.498 14.704 6.549 14.779 / \plot 6.549 14.779 6.598 14.848 / \plot 6.598 14.848 6.648 14.918 / \plot 6.648 14.918 6.699 14.984 / \plot 6.699 14.984 6.750 15.050 / \plot 6.750 15.050 6.801 15.111 / \plot 6.801 15.111 6.854 15.172 / \plot 6.854 15.172 6.905 15.234 / \plot 6.905 15.234 6.960 15.293 / \plot 6.960 15.293 7.015 15.354 / \plot 7.015 15.354 7.074 15.416 / \plot 7.074 15.416 7.133 15.475 / \plot 7.133 15.475 7.195 15.534 / \plot 7.195 15.534 7.256 15.593 / \plot 7.256 15.593 7.319 15.653 / \plot 7.319 15.653 7.385 15.710 / \plot 7.385 15.710 7.453 15.767 / \plot 7.453 15.767 7.521 15.824 / \plot 7.521 15.824 7.592 15.881 / \plot 7.592 15.881 7.664 15.936 / \plot 7.664 15.936 7.739 15.991 / \plot 7.739 15.991 7.815 16.046 / \plot 7.815 16.046 7.891 16.099 / \plot 7.891 16.099 7.969 16.152 / \plot 7.969 16.152 8.050 16.203 / \plot 8.050 16.203 8.132 16.254 / \plot 8.132 16.254 8.215 16.303 / \plot 8.215 16.303 8.297 16.351 / \plot 8.297 16.351 8.382 16.396 / \plot 8.382 16.396 8.467 16.440 / \plot 8.467 16.440 8.553 16.482 / \plot 8.553 16.482 8.638 16.525 / \plot 8.638 16.525 8.725 16.563 / \plot 8.725 16.563 8.812 16.601 / \plot 8.812 16.601 8.896 16.635 / \plot 8.896 16.635 8.983 16.669 / \plot 8.983 16.669 9.068 16.698 / \plot 9.068 16.698 9.155 16.728 / \plot 9.155 16.728 9.239 16.753 / \plot 9.239 16.753 9.322 16.779 / \plot 9.322 16.779 9.406 16.802 / \plot 9.406 16.802 9.489 16.821 / \plot 9.489 16.821 9.572 16.840 / \plot 9.572 16.840 9.652 16.855 / \plot 9.652 16.855 9.732 16.870 / \plot 9.732 16.870 9.815 16.883 / \plot 9.815 16.883 9.893 16.893 / \plot 9.893 16.893 9.974 16.899 / \plot 9.974 16.899 10.054 16.906 / \plot 10.054 16.906 10.139 16.912 / \plot 10.139 16.912 10.224 16.914 / \putrule from 10.224 16.914 to 10.308 16.914 \putrule from 10.308 16.914 to 10.395 16.914 \plot 10.395 16.914 10.482 16.910 / \plot 10.482 16.910 10.569 16.906 / \plot 10.569 16.906 10.657 16.899 / \plot 10.657 16.899 10.746 16.889 / \plot 10.746 16.889 10.837 16.878 / \plot 10.837 16.878 10.928 16.866 / \plot 10.928 16.866 11.019 16.851 / \plot 11.019 16.851 11.110 16.836 / \plot 11.110 16.836 11.204 16.817 / \plot 11.204 16.817 11.295 16.798 / \plot 11.295 16.798 11.388 16.777 / \plot 11.388 16.777 11.479 16.753 / \plot 11.479 16.753 11.570 16.730 / \plot 11.570 16.730 11.661 16.705 / \plot 11.661 16.705 11.750 16.679 / \plot 11.750 16.679 11.839 16.652 / \plot 11.839 16.652 11.925 16.622 / \plot 11.925 16.622 12.010 16.593 / \plot 12.010 16.593 12.093 16.563 / \plot 12.093 16.563 12.175 16.531 / \plot 12.175 16.531 12.253 16.499 / \plot 12.253 16.499 12.330 16.468 / \plot 12.330 16.468 12.404 16.436 / \plot 12.404 16.436 12.476 16.404 / \plot 12.476 16.404 12.545 16.372 / \plot 12.545 16.372 12.611 16.339 / \plot 12.611 16.339 12.675 16.307 / \plot 12.675 16.307 12.736 16.275 / \plot 12.736 16.275 12.793 16.243 / \plot 12.793 16.243 12.848 16.212 / \plot 12.848 16.212 12.901 16.180 / \plot 12.901 16.180 12.952 16.148 / \plot 12.952 16.148 12.998 16.118 / \plot 12.998 16.118 13.045 16.087 / \plot 13.045 16.087 13.083 16.059 / \plot 13.083 16.059 13.119 16.032 / \plot 13.119 16.032 13.155 16.004 / \plot 13.155 16.004 13.189 15.977 / \plot 13.189 15.977 13.221 15.947 / \plot 13.221 15.947 13.250 15.919 / \plot 13.250 15.919 13.278 15.890 / \plot 13.278 15.890 13.303 15.860 / \plot 13.303 15.860 13.327 15.828 / \plot 13.327 15.828 13.348 15.799 / \plot 13.348 15.799 13.367 15.767 / \plot 13.367 15.767 13.382 15.733 / \plot 13.382 15.733 13.396 15.699 / \plot 13.396 15.699 13.407 15.665 / \plot 13.407 15.665 13.415 15.629 / \plot 13.415 15.629 13.420 15.591 / \plot 13.420 15.591 13.422 15.553 / \putrule from 13.422 15.553 to 13.422 15.515 \plot 13.422 15.515 13.418 15.475 / \plot 13.418 15.475 13.409 15.433 / \plot 13.409 15.433 13.399 15.390 / \plot 13.399 15.390 13.386 15.346 / \plot 13.386 15.346 13.369 15.301 / \plot 13.369 15.301 13.348 15.255 / \plot 13.348 15.255 13.324 15.208 / \plot 13.324 15.208 13.299 15.160 / \plot 13.299 15.160 13.267 15.111 / \plot 13.267 15.111 13.233 15.060 / \plot 13.233 15.060 13.197 15.009 / \plot 13.197 15.009 13.157 14.956 / \plot 13.157 14.956 13.115 14.903 / \plot 13.115 14.903 13.068 14.848 / \plot 13.068 14.848 13.020 14.793 / \plot 13.020 14.793 12.967 14.736 / \plot 12.967 14.736 12.912 14.679 / \plot 12.912 14.679 12.852 14.620 / \plot 12.852 14.620 12.791 14.558 / \plot 12.791 14.558 12.728 14.497 / \plot 12.728 14.497 12.660 14.434 / \plot 12.660 14.434 12.588 14.370 / \plot 12.588 14.370 12.514 14.302 / \plot 12.514 14.302 12.435 14.235 / \plot 12.435 14.235 12.370 14.177 / \plot 12.370 14.177 12.302 14.120 / \plot 12.302 14.120 12.232 14.061 / \plot 12.232 14.061 12.158 14.000 / \plot 12.158 14.000 12.082 13.936 / \plot 12.082 13.936 12.004 13.873 / \plot 12.004 13.873 11.923 13.807 / \plot 11.923 13.807 11.841 13.741 / \plot 11.841 13.741 11.754 13.672 / \plot 11.754 13.672 11.665 13.602 / \plot 11.665 13.602 11.574 13.532 / \plot 11.574 13.532 11.481 13.458 / \plot 11.481 13.458 11.386 13.386 / \plot 11.386 13.386 11.286 13.310 / \plot 11.286 13.310 11.187 13.233 / \plot 11.187 13.233 11.083 13.157 / \plot 11.083 13.157 10.979 13.079 / \plot 10.979 13.079 10.871 13.001 / \plot 10.871 13.001 10.763 12.920 / \plot 10.763 12.920 10.653 12.840 / \plot 10.653 12.840 10.543 12.759 / \plot 10.543 12.759 10.431 12.679 / \plot 10.431 12.679 10.317 12.596 / \plot 10.317 12.596 10.202 12.516 / \plot 10.202 12.516 10.088 12.435 / \plot 10.088 12.435 9.972 12.355 / \plot 9.972 12.355 9.855 12.275 / \plot 9.855 12.275 9.741 12.194 / \plot 9.741 12.194 9.624 12.116 / \plot 9.624 12.116 9.510 12.037 / \plot 9.510 12.037 9.396 11.961 / \plot 9.396 11.961 9.282 11.885 / \plot 9.282 11.885 9.167 11.811 / \plot 9.167 11.811 9.057 11.737 / \plot 9.057 11.737 8.945 11.665 / \plot 8.945 11.665 8.837 11.595 / \plot 8.837 11.595 8.729 11.527 / \plot 8.729 11.527 8.623 11.460 / \plot 8.623 11.460 8.520 11.394 / \plot 8.520 11.394 8.416 11.333 / \plot 8.416 11.333 8.316 11.271 / \plot 8.316 11.271 8.217 11.212 / \plot 8.217 11.212 8.122 11.157 / \plot 8.122 11.157 8.029 11.102 / \plot 8.029 11.102 7.935 11.049 / \plot 7.935 11.049 7.846 10.998 / \plot 7.846 10.998 7.760 10.950 / \plot 7.760 10.950 7.675 10.905 / \plot 7.675 10.905 7.590 10.861 / \plot 7.590 10.861 7.510 10.818 / \plot 7.510 10.818 7.432 10.780 / \plot 7.432 10.780 7.355 10.742 / \plot 7.355 10.742 7.243 10.689 / \plot 7.243 10.689 7.133 10.640 / \plot 7.133 10.640 7.029 10.598 / \plot 7.029 10.598 6.930 10.560 / \plot 6.930 10.560 6.833 10.528 / \plot 6.833 10.528 6.739 10.501 / \plot 6.739 10.501 6.648 10.478 / \plot 6.648 10.478 6.560 10.458 / \plot 6.560 10.458 6.475 10.446 / \plot 6.475 10.446 6.388 10.437 / \plot 6.388 10.437 6.303 10.433 / \putrule from 6.303 10.433 to 6.221 10.433 \plot 6.221 10.433 6.136 10.437 / \plot 6.136 10.437 6.054 10.446 / \plot 6.054 10.446 5.971 10.458 / \plot 5.971 10.458 5.889 10.473 / \plot 5.889 10.473 5.806 10.492 / \plot 5.806 10.492 5.726 10.516 / \plot 5.726 10.516 5.647 10.541 / \plot 5.647 10.541 5.569 10.566 / \plot 5.569 10.566 5.495 10.594 / \plot 5.495 10.594 5.425 10.624 / \plot 5.425 10.624 5.362 10.651 / \plot 5.362 10.651 5.302 10.679 / \plot 5.302 10.679 5.249 10.704 / \plot 5.249 10.704 5.205 10.727 / \plot 5.205 10.727 5.167 10.746 / \plot 5.167 10.746 5.137 10.763 / \plot 5.137 10.763 5.114 10.776 / \plot 5.114 10.776 5.097 10.784 / \plot 5.097 10.784 5.088 10.791 / \plot 5.088 10.791 5.082 10.793 / \plot 5.082 10.793 5.080 10.795 / }% \linethickness= 0.500pt \setplotsymbol ({\thinlinefont .}) {\color[rgb]{0,0,0}\plot 5.080 12.065 5.082 12.069 / \plot 5.082 12.069 5.086 12.076 / \plot 5.086 12.076 5.095 12.090 / \plot 5.095 12.090 5.108 12.112 / \plot 5.108 12.112 5.127 12.143 / \plot 5.127 12.143 5.150 12.181 / \plot 5.150 12.181 5.177 12.230 / \plot 5.177 12.230 5.211 12.285 / \plot 5.211 12.285 5.249 12.349 / \plot 5.249 12.349 5.292 12.416 / \plot 5.292 12.416 5.334 12.488 / \plot 5.334 12.488 5.381 12.565 / \plot 5.381 12.565 5.429 12.639 / \plot 5.429 12.639 5.476 12.715 / \plot 5.476 12.715 5.524 12.789 / \plot 5.524 12.789 5.571 12.863 / \plot 5.571 12.863 5.618 12.933 / \plot 5.618 12.933 5.664 13.003 / \plot 5.664 13.003 5.709 13.068 / \plot 5.709 13.068 5.753 13.132 / \plot 5.753 13.132 5.798 13.193 / \plot 5.798 13.193 5.840 13.252 / \plot 5.840 13.252 5.884 13.312 / \plot 5.884 13.312 5.929 13.369 / \plot 5.929 13.369 5.973 13.424 / \plot 5.973 13.424 6.020 13.481 / \plot 6.020 13.481 6.066 13.538 / \plot 6.066 13.538 6.115 13.595 / \plot 6.115 13.595 6.166 13.652 / \plot 6.166 13.652 6.208 13.701 / \plot 6.208 13.701 6.253 13.752 / \plot 6.253 13.752 6.297 13.801 / \plot 6.297 13.801 6.346 13.854 / \plot 6.346 13.854 6.394 13.906 / \plot 6.394 13.906 6.445 13.959 / \plot 6.445 13.959 6.498 14.014 / \plot 6.498 14.014 6.551 14.069 / \plot 6.551 14.069 6.608 14.127 / \plot 6.608 14.127 6.665 14.186 / \plot 6.665 14.186 6.725 14.243 / \plot 6.725 14.243 6.786 14.304 / \plot 6.786 14.304 6.847 14.364 / \plot 6.847 14.364 6.911 14.425 / \plot 6.911 14.425 6.977 14.486 / \plot 6.977 14.486 7.042 14.550 / \plot 7.042 14.550 7.110 14.611 / \plot 7.110 14.611 7.178 14.673 / \plot 7.178 14.673 7.245 14.736 / \plot 7.245 14.736 7.313 14.798 / \plot 7.313 14.798 7.383 14.857 / \plot 7.383 14.857 7.451 14.918 / \plot 7.451 14.918 7.521 14.978 / \plot 7.521 14.978 7.588 15.037 / \plot 7.588 15.037 7.656 15.094 / \plot 7.656 15.094 7.724 15.149 / \plot 7.724 15.149 7.791 15.204 / \plot 7.791 15.204 7.857 15.257 / \plot 7.857 15.257 7.923 15.310 / \plot 7.923 15.310 7.988 15.361 / \plot 7.988 15.361 8.052 15.409 / \plot 8.052 15.409 8.115 15.458 / \plot 8.115 15.458 8.177 15.505 / \plot 8.177 15.505 8.238 15.549 / \plot 8.238 15.549 8.299 15.593 / \plot 8.299 15.593 8.361 15.636 / \plot 8.361 15.636 8.424 15.680 / \plot 8.424 15.680 8.490 15.725 / \plot 8.490 15.725 8.553 15.767 / \plot 8.553 15.767 8.619 15.809 / \plot 8.619 15.809 8.683 15.852 / \plot 8.683 15.852 8.750 15.892 / \plot 8.750 15.892 8.816 15.932 / \plot 8.816 15.932 8.884 15.970 / \plot 8.884 15.970 8.951 16.010 / \plot 8.951 16.010 9.019 16.049 / \plot 9.019 16.049 9.089 16.087 / \plot 9.089 16.087 9.159 16.123 / \plot 9.159 16.123 9.231 16.161 / \plot 9.231 16.161 9.301 16.195 / \plot 9.301 16.195 9.373 16.231 / \plot 9.373 16.231 9.442 16.264 / \plot 9.442 16.264 9.514 16.296 / \plot 9.514 16.296 9.584 16.328 / \plot 9.584 16.328 9.656 16.358 / \plot 9.656 16.358 9.726 16.387 / \plot 9.726 16.387 9.794 16.415 / \plot 9.794 16.415 9.864 16.442 / \plot 9.864 16.442 9.931 16.468 / \plot 9.931 16.468 9.997 16.491 / \plot 9.997 16.491 10.063 16.514 / \plot 10.063 16.514 10.128 16.535 / \plot 10.128 16.535 10.192 16.557 / \plot 10.192 16.557 10.253 16.576 / \plot 10.253 16.576 10.315 16.595 / \plot 10.315 16.595 10.376 16.612 / \plot 10.376 16.612 10.435 16.626 / \plot 10.435 16.626 10.494 16.641 / \plot 10.494 16.641 10.552 16.656 / \plot 10.552 16.656 10.611 16.669 / \plot 10.611 16.669 10.674 16.684 / \plot 10.674 16.684 10.740 16.696 / \plot 10.740 16.696 10.808 16.709 / \plot 10.808 16.709 10.873 16.720 / \plot 10.873 16.720 10.941 16.730 / \plot 10.941 16.730 11.009 16.739 / \plot 11.009 16.739 11.079 16.747 / \plot 11.079 16.747 11.151 16.753 / \plot 11.151 16.753 11.220 16.760 / \plot 11.220 16.760 11.295 16.764 / \plot 11.295 16.764 11.366 16.768 / \putrule from 11.366 16.768 to 11.441 16.768 \plot 11.441 16.768 11.515 16.770 / \plot 11.515 16.770 11.589 16.768 / \plot 11.589 16.768 11.665 16.766 / \plot 11.665 16.766 11.739 16.760 / \plot 11.739 16.760 11.813 16.756 / \plot 11.813 16.756 11.887 16.747 / \plot 11.887 16.747 11.961 16.739 / \plot 11.961 16.739 12.033 16.726 / \plot 12.033 16.726 12.105 16.715 / \plot 12.105 16.715 12.175 16.701 / \plot 12.175 16.701 12.245 16.686 / \plot 12.245 16.686 12.313 16.669 / \plot 12.313 16.669 12.378 16.650 / \plot 12.378 16.650 12.444 16.631 / \plot 12.444 16.631 12.509 16.609 / \plot 12.509 16.609 12.573 16.586 / \plot 12.573 16.586 12.636 16.563 / \plot 12.636 16.563 12.700 16.535 / \plot 12.700 16.535 12.747 16.516 / \plot 12.747 16.516 12.795 16.495 / \plot 12.795 16.495 12.842 16.472 / \plot 12.842 16.472 12.891 16.447 / \plot 12.891 16.447 12.939 16.421 / \plot 12.939 16.421 12.988 16.394 / \plot 12.988 16.394 13.037 16.364 / \plot 13.037 16.364 13.085 16.332 / \plot 13.085 16.332 13.136 16.298 / \plot 13.136 16.298 13.185 16.264 / \plot 13.185 16.264 13.236 16.226 / \plot 13.236 16.226 13.286 16.184 / \plot 13.286 16.184 13.339 16.142 / \plot 13.339 16.142 13.390 16.097 / \plot 13.390 16.097 13.441 16.049 / \plot 13.441 16.049 13.492 15.998 / \plot 13.492 15.998 13.542 15.945 / \plot 13.542 15.945 13.593 15.888 / \plot 13.593 15.888 13.644 15.828 / \plot 13.644 15.828 13.695 15.767 / \plot 13.695 15.767 13.746 15.704 / \plot 13.746 15.704 13.794 15.636 / \plot 13.794 15.636 13.843 15.566 / \plot 13.843 15.566 13.890 15.494 / \plot 13.890 15.494 13.936 15.420 / \plot 13.936 15.420 13.983 15.344 / \plot 13.983 15.344 14.027 15.263 / \plot 14.027 15.263 14.069 15.183 / \plot 14.069 15.183 14.112 15.098 / \plot 14.112 15.098 14.154 15.014 / \plot 14.154 15.014 14.194 14.925 / \plot 14.194 14.925 14.232 14.834 / \plot 14.232 14.834 14.271 14.740 / \plot 14.271 14.740 14.307 14.645 / \plot 14.307 14.645 14.340 14.548 / \plot 14.340 14.548 14.374 14.448 / \plot 14.374 14.448 14.408 14.347 / \plot 14.408 14.347 14.440 14.241 / \plot 14.440 14.241 14.470 14.133 / \plot 14.470 14.133 14.499 14.023 / \plot 14.499 14.023 14.520 13.942 / \plot 14.520 13.942 14.539 13.858 / \plot 14.539 13.858 14.561 13.773 / \plot 14.561 13.773 14.580 13.684 / \plot 14.580 13.684 14.599 13.593 / \plot 14.599 13.593 14.616 13.500 / \plot 14.616 13.500 14.635 13.405 / \plot 14.635 13.405 14.652 13.303 / \plot 14.652 13.303 14.669 13.202 / \plot 14.669 13.202 14.688 13.094 / \plot 14.688 13.094 14.704 12.984 / \plot 14.704 12.984 14.721 12.867 / \plot 14.721 12.867 14.738 12.749 / \plot 14.738 12.749 14.753 12.624 / \plot 14.753 12.624 14.770 12.495 / \plot 14.770 12.495 14.787 12.359 / \plot 14.787 12.359 14.804 12.220 / \plot 14.804 12.220 14.821 12.076 / \plot 14.821 12.076 14.836 11.925 / \plot 14.836 11.925 14.853 11.769 / \plot 14.853 11.769 14.870 11.608 / \plot 14.870 11.608 14.887 11.441 / \plot 14.887 11.441 14.903 11.267 / \plot 14.903 11.267 14.920 11.089 / \plot 14.920 11.089 14.937 10.907 / \plot 14.937 10.907 14.954 10.719 / \plot 14.954 10.719 14.969 10.526 / \plot 14.969 10.526 14.986 10.331 / \plot 14.986 10.331 15.003 10.132 / \plot 15.003 10.132 15.020 9.931 / \plot 15.020 9.931 15.037 9.728 / \plot 15.037 9.728 15.054 9.525 / \plot 15.054 9.525 15.069 9.322 / \plot 15.069 9.322 15.085 9.121 / \plot 15.085 9.121 15.100 8.924 / \plot 15.100 8.924 15.115 8.729 / \plot 15.115 8.729 15.128 8.541 / \plot 15.128 8.541 15.143 8.361 / \plot 15.143 8.361 15.155 8.187 / \plot 15.155 8.187 15.166 8.022 / \plot 15.166 8.022 15.179 7.870 / \plot 15.179 7.870 15.189 7.728 / \plot 15.189 7.728 15.198 7.599 / \plot 15.198 7.599 15.206 7.482 / \plot 15.206 7.482 15.212 7.379 / \plot 15.212 7.379 15.219 7.290 / \plot 15.219 7.290 15.225 7.211 / \plot 15.225 7.211 15.229 7.148 / \plot 15.229 7.148 15.232 7.097 / \plot 15.232 7.097 15.236 7.057 / \plot 15.236 7.057 15.238 7.027 / \putrule from 15.238 7.027 to 15.238 7.008 \plot 15.238 7.008 15.240 6.996 / \putrule from 15.240 6.996 to 15.240 6.987 \putrule from 15.240 6.987 to 15.240 6.985 }% \put{\SetFigFont{6}{7.2}{\rmdefault}{\mddefault}{\updefault}{\color[rgb]{0,0,0}$\bar{N}_-$}% } [lB] at 3.810 10.636 \put{\SetFigFont{6}{7.2}{\rmdefault}{\mddefault}{\updefault}{\color[rgb]{0,0,0}$\tilde{S}_-$}% } [lB] at 3.810 12.065 \put{\SetFigFont{6}{7.2}{\rmdefault}{\mddefault}{\updefault}{\color[rgb]{0,0,0}$\tilde{N}_-$}% } [lB] at 3.810 20.796 \put{\SetFigFont{6}{7.2}{\rmdefault}{\mddefault}{\updefault}{\color[rgb]{0,0,0}$\tilde{N}_+$}% } [lB] at 15.399 16.986 \put{\SetFigFont{6}{7.2}{\rmdefault}{\mddefault}{\updefault}{\color[rgb]{0,0,0}$\tilde{S}_+$}% } [lB] at 15.399 15.716 \put{\SetFigFont{6}{7.2}{\rmdefault}{\mddefault}{\updefault}{\color[rgb]{0,0,0}$\bar{N}_+$}% } [lB] at 15.399 6.826 \put{\SetFigFont{6}{7.2}{\rmdefault}{\mddefault}{\updefault}{\color[rgb]{0,0,0}$L$}% } [lB] at 7.144 20.003 \put{\SetFigFont{6}{7.2}{\rmdefault}{\mddefault}{\updefault}{\color[rgb]{0,0,0}$L'$}% } [lB] at 12.700 18.415 \put{\SetFigFont{6}{7.2}{\rmdefault}{\mddefault}{\updefault}{\color[rgb]{0,0,0}$\mathcal{W}^-$}% } [lB] at 7.303 14.605 \put{\SetFigFont{6}{7.2}{\rmdefault}{\mddefault}{\updefault}{\color[rgb]{0,0,0}$\Gamma_l$}% } [lB] at 11.271 13.018 \linethickness=0pt \putrectangle corners at 3.778 21.660 and 15.431 6.280 \endpicture} \end{center} \caption{Construction of a barrier} \end{figure} Let $\Upsilon$ be the curve in $\mathcal{C}$ consisting of two line segments: From $(-\frac{\pi}{2},\frac{\pi}{2} - \eta_0)$ along the horizontal line $L'$, upto the intersection point of $L$ and $L'$, and then from there along $L$ to $\widetilde{S}_+$. The curve $\Upsilon$ provides a barrier for the flow: no orbit can cross it from the region below $\Upsilon$ into the region above $\Upsilon$. Let $\widetilde{\mathcal{W}}^-$ be the unstable manifold af $\widetilde{S}_-$. Thus $\widetilde{\mathcal{W}}^-$ must stay below $\Upsilon$, and as a result the $\omega$-limit of $\mathcal{W}^-$ cannot be $\widetilde{N}_+$ so therefore its winding number is not zero or negative. \end{proof} \begin{prop}\label{prop:om0} Given $\lambda \leq -1+a$ and $E\in[0,E_l(\lambda))$ the corridor ${\mathcal K}_1(E,\lambda)$ of the flow \refeq{dynsysOm} has winding number equal to zero. \end{prop} \begin{proof} Once again we find a barrier that prevents $\mathcal{W}^-$ from going down: Let us compute the slope of orbits crossing the line $$ L = \{(\xi,\Omega)\ |\ \Omega = \xi-{\textstyle\frac{\pi}{2}}\} $$ and compare it to the slope of this line. $$ j(\xi):= g_{E,\lambda}(\xi,\xi-{\textstyle\frac{\pi}{2}}) - \cos^2\xi = 2a(1-E) -2(\lambda+a)\cos^2\xi + 2\gamma\sin\xi\cos\xi. $$ Thus $j(\pm\pi/2) = 2a(1-E)>0$. Any interior minimum of $j$ must be achieved at a critical point: $$ j'(\xi_0) = 0 \implies \xi_0 = {\textstyle\frac{1}{2}} \tan^{-1}\frac{\gamma}{\lambda+a} >0. $$ However we have $$ j(\xi_0) = 2a(1-E) -(\lambda+a)- (\lambda+a)\cos2\xi_0 +\gamma\sin2\xi_0 = 2a(1-E)- \frac{2(\lambda+a)^3}{(\lambda+a)^2 +\gamma^2}> 2a(1-E). $$ \begin{figure}[ht] \begin{center} \font\thinlinefont=cmr5 \begingroup\makeatletter\ifx\SetFigFont\undefined% \gdef\SetFigFont#1#2#3#4#5{% \reset@font\fontsize{#1}{#2pt}% \fontfamily{#3}\fontseries{#4}\fontshape{#5}% \selectfont}% \fi\endgroup% \mbox{\beginpicture \setcoordinatesystem units <0.50000cm,0.50000cm> \unitlength=0.50000cm \linethickness=1pt \setplotsymbol ({\makebox(0,0)[l]{\tencirc\symbol{'160}}}) \setshadesymbol ({\thinlinefont .}) \setlinear \linethickness= 0.500pt \setplotsymbol ({\thinlinefont .}) {\color[rgb]{0,0,0}\putrule from 5.080 11.430 to 15.240 11.430 }% \linethickness=1pt \setplotsymbol ({\makebox(0,0)[l]{\tencirc\symbol{'160}}}) {\color[rgb]{0,0,0}\putrule from 5.080 21.590 to 15.240 21.590 }% \linethickness= 0.500pt \setplotsymbol ({\thinlinefont .}) {\color[rgb]{0,0,0}\putrule from 5.080 17.462 to 5.080 16.034 \plot 5.017 16.288 5.080 16.034 5.143 16.288 / }% \linethickness= 0.500pt \setplotsymbol ({\thinlinefont .}) {\color[rgb]{0,0,0}\putrule from 15.240 12.383 to 15.240 10.795 \plot 15.176 11.049 15.240 10.795 15.304 11.049 / }% \linethickness= 0.500pt \setplotsymbol ({\thinlinefont .}) {\color[rgb]{0,0,0}\putrule from 5.080 16.510 to 15.240 16.510 }% \linethickness=1pt \setplotsymbol ({\makebox(0,0)[l]{\tencirc\symbol{'160}}}) {\color[rgb]{0,0,0}\putrule from 15.240 21.590 to 15.240 6.315 \putrule from 15.240 6.350 to 5.045 6.350 \putrule from 5.080 6.350 to 5.080 21.590 }% \linethickness= 0.500pt \setplotsymbol ({\thinlinefont .}) {\color[rgb]{0,0,0}\plot 5.080 11.430 15.240 16.510 / }% \linethickness= 0.500pt \setplotsymbol ({\thinlinefont .}) {\color[rgb]{0,0,0}\plot 7.620 12.383 8.255 13.335 / \plot 8.167 13.088 8.255 13.335 8.061 13.159 / }% \linethickness= 0.500pt \setplotsymbol ({\thinlinefont .}) {\color[rgb]{0,0,0}\plot 12.065 14.605 12.700 15.558 / \plot 12.612 15.311 12.700 15.558 12.506 15.381 / }% \linethickness= 0.500pt \setplotsymbol ({\thinlinefont .}) {\color[rgb]{0,0,0}\plot 15.240 17.462 15.236 17.458 / \plot 15.236 17.458 15.225 17.452 / \plot 15.225 17.452 15.208 17.439 / \plot 15.208 17.439 15.181 17.420 / \plot 15.181 17.420 15.143 17.393 / \plot 15.143 17.393 15.096 17.361 / \plot 15.096 17.361 15.041 17.323 / \plot 15.041 17.323 14.980 17.283 / \plot 14.980 17.283 14.912 17.240 / \plot 14.912 17.240 14.844 17.198 / \plot 14.844 17.198 14.774 17.158 / \plot 14.774 17.158 14.707 17.120 / \plot 14.707 17.120 14.637 17.081 / \plot 14.637 17.081 14.571 17.050 / \plot 14.571 17.050 14.506 17.020 / \plot 14.506 17.020 14.442 16.995 / \plot 14.442 16.995 14.381 16.974 / \plot 14.381 16.974 14.319 16.954 / \plot 14.319 16.954 14.258 16.940 / \plot 14.258 16.940 14.194 16.927 / \plot 14.194 16.927 14.131 16.919 / \plot 14.131 16.919 14.065 16.910 / \plot 14.065 16.910 13.998 16.906 / \putrule from 13.998 16.906 to 13.949 16.906 \plot 13.949 16.906 13.900 16.904 / \plot 13.900 16.904 13.851 16.906 / \putrule from 13.851 16.906 to 13.799 16.906 \plot 13.799 16.906 13.746 16.908 / \plot 13.746 16.908 13.688 16.910 / \plot 13.688 16.910 13.631 16.912 / \plot 13.631 16.912 13.570 16.916 / \plot 13.570 16.916 13.506 16.921 / \plot 13.506 16.921 13.443 16.923 / \plot 13.443 16.923 13.375 16.929 / \plot 13.375 16.929 13.305 16.933 / \plot 13.305 16.933 13.231 16.938 / \plot 13.231 16.938 13.157 16.942 / \plot 13.157 16.942 13.079 16.946 / \plot 13.079 16.946 13.001 16.950 / \plot 13.001 16.950 12.918 16.954 / \plot 12.918 16.954 12.835 16.957 / \plot 12.835 16.957 12.751 16.959 / \plot 12.751 16.959 12.664 16.963 / \putrule from 12.664 16.963 to 12.575 16.963 \plot 12.575 16.963 12.486 16.965 / \putrule from 12.486 16.965 to 12.395 16.965 \plot 12.395 16.965 12.302 16.963 / \plot 12.302 16.963 12.209 16.961 / \plot 12.209 16.961 12.116 16.959 / \plot 12.116 16.959 12.021 16.954 / \plot 12.021 16.954 11.925 16.948 / \plot 11.925 16.948 11.830 16.942 / \plot 11.830 16.942 11.735 16.933 / \plot 11.735 16.933 11.637 16.925 / \plot 11.637 16.925 11.540 16.912 / \plot 11.540 16.912 11.441 16.902 / \plot 11.441 16.902 11.341 16.887 / \plot 11.341 16.887 11.242 16.872 / \plot 11.242 16.872 11.140 16.853 / \plot 11.140 16.853 11.057 16.840 / \plot 11.057 16.840 10.975 16.823 / \plot 10.975 16.823 10.890 16.806 / \plot 10.890 16.806 10.806 16.787 / \plot 10.806 16.787 10.719 16.768 / \plot 10.719 16.768 10.630 16.747 / \plot 10.630 16.747 10.541 16.724 / \plot 10.541 16.724 10.448 16.701 / \plot 10.448 16.701 10.355 16.677 / \plot 10.355 16.677 10.262 16.650 / \plot 10.262 16.650 10.164 16.622 / \plot 10.164 16.622 10.067 16.593 / \plot 10.067 16.593 9.967 16.563 / \plot 9.967 16.563 9.866 16.531 / \plot 9.866 16.531 9.764 16.497 / \plot 9.764 16.497 9.663 16.463 / \plot 9.663 16.463 9.559 16.427 / \plot 9.559 16.427 9.453 16.391 / \plot 9.453 16.391 9.349 16.353 / \plot 9.349 16.353 9.243 16.315 / \plot 9.243 16.315 9.138 16.275 / \plot 9.138 16.275 9.032 16.233 / \plot 9.032 16.233 8.926 16.190 / \plot 8.926 16.190 8.822 16.148 / \plot 8.822 16.148 8.716 16.104 / \plot 8.716 16.104 8.613 16.059 / \plot 8.613 16.059 8.511 16.015 / \plot 8.511 16.015 8.410 15.968 / \plot 8.410 15.968 8.310 15.924 / \plot 8.310 15.924 8.211 15.877 / \plot 8.211 15.877 8.113 15.828 / \plot 8.113 15.828 8.018 15.782 / \plot 8.018 15.782 7.925 15.735 / \plot 7.925 15.735 7.834 15.687 / \plot 7.834 15.687 7.745 15.638 / \plot 7.745 15.638 7.658 15.591 / \plot 7.658 15.591 7.571 15.543 / \plot 7.571 15.543 7.489 15.494 / \plot 7.489 15.494 7.408 15.447 / \plot 7.408 15.447 7.330 15.399 / \plot 7.330 15.399 7.254 15.350 / \plot 7.254 15.350 7.180 15.301 / \plot 7.180 15.301 7.108 15.253 / \plot 7.108 15.253 7.040 15.206 / \plot 7.040 15.206 6.972 15.157 / \plot 6.972 15.157 6.907 15.107 / \plot 6.907 15.107 6.828 15.047 / \plot 6.828 15.047 6.754 14.988 / \plot 6.754 14.988 6.682 14.927 / \plot 6.682 14.927 6.612 14.863 / \plot 6.612 14.863 6.545 14.800 / \plot 6.545 14.800 6.479 14.734 / \plot 6.479 14.734 6.416 14.666 / \plot 6.416 14.666 6.352 14.597 / \plot 6.352 14.597 6.291 14.525 / \plot 6.291 14.525 6.229 14.448 / \plot 6.229 14.448 6.168 14.368 / \plot 6.168 14.368 6.109 14.285 / \plot 6.109 14.285 6.049 14.201 / \plot 6.049 14.201 5.988 14.110 / \plot 5.988 14.110 5.929 14.017 / \plot 5.929 14.017 5.867 13.917 / \plot 5.867 13.917 5.808 13.818 / \plot 5.808 13.818 5.747 13.712 / \plot 5.747 13.712 5.687 13.606 / \plot 5.687 13.606 5.628 13.496 / \plot 5.628 13.496 5.569 13.386 / \plot 5.569 13.386 5.512 13.276 / \plot 5.512 13.276 5.455 13.166 / \plot 5.455 13.166 5.402 13.060 / \plot 5.402 13.060 5.351 12.956 / \plot 5.351 12.956 5.304 12.861 / \plot 5.304 12.861 5.260 12.770 / \plot 5.260 12.770 5.222 12.687 / \plot 5.222 12.687 5.188 12.615 / \plot 5.188 12.615 5.158 12.554 / \plot 5.158 12.554 5.133 12.501 / \plot 5.133 12.501 5.116 12.461 / \plot 5.116 12.461 5.101 12.429 / \plot 5.101 12.429 5.091 12.408 / \plot 5.091 12.408 5.084 12.393 / \plot 5.084 12.393 5.082 12.387 / \plot 5.082 12.387 5.080 12.383 / }% \linethickness= 0.500pt \setplotsymbol ({\thinlinefont .}) {\color[rgb]{0,0,0}\plot 15.240 15.558 15.236 15.555 / \plot 15.236 15.555 15.225 15.553 / \plot 15.225 15.553 15.206 15.547 / \plot 15.206 15.547 15.179 15.538 / \plot 15.179 15.538 15.141 15.526 / \plot 15.141 15.526 15.092 15.511 / \plot 15.092 15.511 15.035 15.492 / \plot 15.035 15.492 14.971 15.471 / \plot 14.971 15.471 14.901 15.447 / \plot 14.901 15.447 14.827 15.420 / \plot 14.827 15.420 14.753 15.392 / \plot 14.753 15.392 14.677 15.365 / \plot 14.677 15.365 14.605 15.335 / \plot 14.605 15.335 14.533 15.306 / \plot 14.533 15.306 14.465 15.276 / \plot 14.465 15.276 14.402 15.246 / \plot 14.402 15.246 14.338 15.215 / \plot 14.338 15.215 14.281 15.183 / \plot 14.281 15.183 14.224 15.151 / \plot 14.224 15.151 14.171 15.117 / \plot 14.171 15.117 14.118 15.083 / \plot 14.118 15.083 14.067 15.045 / \plot 14.067 15.045 14.017 15.007 / \plot 14.017 15.007 13.968 14.967 / \plot 13.968 14.967 13.917 14.922 / \plot 13.917 14.922 13.879 14.887 / \plot 13.879 14.887 13.839 14.851 / \plot 13.839 14.851 13.801 14.812 / \plot 13.801 14.812 13.758 14.772 / \plot 13.758 14.772 13.718 14.730 / \plot 13.718 14.730 13.676 14.685 / \plot 13.676 14.685 13.633 14.641 / \plot 13.633 14.641 13.589 14.592 / \plot 13.589 14.592 13.542 14.541 / \plot 13.542 14.541 13.496 14.491 / \plot 13.496 14.491 13.449 14.436 / \plot 13.449 14.436 13.401 14.379 / \plot 13.401 14.379 13.352 14.321 / \plot 13.352 14.321 13.301 14.262 / \plot 13.301 14.262 13.248 14.201 / \plot 13.248 14.201 13.197 14.137 / \plot 13.197 14.137 13.142 14.074 / \plot 13.142 14.074 13.089 14.008 / \plot 13.089 14.008 13.034 13.942 / \plot 13.034 13.942 12.979 13.877 / \plot 12.979 13.877 12.924 13.809 / \plot 12.924 13.809 12.869 13.741 / \plot 12.869 13.741 12.814 13.674 / \plot 12.814 13.674 12.759 13.606 / \plot 12.759 13.606 12.702 13.538 / \plot 12.702 13.538 12.647 13.470 / \plot 12.647 13.470 12.592 13.403 / \plot 12.592 13.403 12.535 13.335 / \plot 12.535 13.335 12.480 13.269 / \plot 12.480 13.269 12.425 13.202 / \plot 12.425 13.202 12.368 13.136 / \plot 12.368 13.136 12.311 13.070 / \plot 12.311 13.070 12.256 13.005 / \plot 12.256 13.005 12.198 12.937 / \plot 12.198 12.937 12.148 12.882 / \plot 12.148 12.882 12.097 12.825 / \plot 12.097 12.825 12.046 12.768 / \plot 12.046 12.768 11.995 12.708 / \plot 11.995 12.708 11.940 12.651 / \plot 11.940 12.651 11.885 12.592 / \plot 11.885 12.592 11.830 12.533 / \plot 11.830 12.533 11.773 12.471 / \plot 11.773 12.471 11.714 12.410 / \plot 11.714 12.410 11.652 12.349 / \plot 11.652 12.349 11.591 12.287 / \plot 11.591 12.287 11.527 12.224 / \plot 11.527 12.224 11.462 12.160 / \plot 11.462 12.160 11.394 12.099 / \plot 11.394 12.099 11.326 12.035 / \plot 11.326 12.035 11.256 11.972 / \plot 11.256 11.972 11.184 11.908 / \plot 11.184 11.908 11.113 11.847 / \plot 11.113 11.847 11.038 11.786 / \plot 11.038 11.786 10.964 11.722 / \plot 10.964 11.722 10.888 11.663 / \plot 10.888 11.663 10.812 11.604 / \plot 10.812 11.604 10.734 11.544 / \plot 10.734 11.544 10.655 11.487 / \plot 10.655 11.487 10.577 11.430 / \plot 10.577 11.430 10.497 11.375 / \plot 10.497 11.375 10.418 11.322 / \plot 10.418 11.322 10.338 11.271 / \plot 10.338 11.271 10.257 11.223 / \plot 10.257 11.223 10.179 11.174 / \plot 10.179 11.174 10.099 11.127 / \plot 10.099 11.127 10.018 11.083 / \plot 10.018 11.083 9.938 11.041 / \plot 9.938 11.041 9.857 11.000 / \plot 9.857 11.000 9.775 10.962 / \plot 9.775 10.962 9.694 10.924 / \plot 9.694 10.924 9.614 10.890 / \plot 9.614 10.890 9.531 10.856 / \plot 9.531 10.856 9.449 10.825 / \plot 9.449 10.825 9.366 10.795 / \plot 9.366 10.795 9.290 10.770 / \plot 9.290 10.770 9.212 10.744 / \plot 9.212 10.744 9.133 10.721 / \plot 9.133 10.721 9.053 10.700 / \plot 9.053 10.700 8.970 10.679 / \plot 8.970 10.679 8.886 10.660 / \plot 8.886 10.660 8.799 10.643 / \plot 8.799 10.643 8.710 10.626 / \plot 8.710 10.626 8.617 10.611 / \plot 8.617 10.611 8.520 10.596 / \plot 8.520 10.596 8.420 10.581 / \plot 8.420 10.581 8.316 10.569 / \plot 8.316 10.569 8.208 10.558 / \plot 8.208 10.558 8.096 10.547 / \plot 8.096 10.547 7.978 10.537 / \plot 7.978 10.537 7.857 10.528 / \plot 7.857 10.528 7.730 10.520 / \plot 7.730 10.520 7.599 10.513 / \plot 7.599 10.513 7.463 10.507 / \plot 7.463 10.507 7.324 10.501 / \plot 7.324 10.501 7.180 10.494 / \plot 7.180 10.494 7.032 10.490 / \plot 7.032 10.490 6.881 10.486 / \plot 6.881 10.486 6.731 10.484 / \plot 6.731 10.484 6.579 10.480 / \plot 6.579 10.480 6.428 10.478 / \putrule from 6.428 10.478 to 6.278 10.478 \plot 6.278 10.478 6.132 10.475 / \putrule from 6.132 10.475 to 5.992 10.475 \plot 5.992 10.475 5.857 10.473 / \putrule from 5.857 10.473 to 5.732 10.473 \putrule from 5.732 10.473 to 5.616 10.473 \putrule from 5.616 10.473 to 5.510 10.473 \plot 5.510 10.473 5.417 10.475 / \putrule from 5.417 10.475 to 5.334 10.475 \putrule from 5.334 10.475 to 5.264 10.475 \putrule from 5.264 10.475 to 5.207 10.475 \plot 5.207 10.475 5.163 10.478 / \putrule from 5.163 10.478 to 5.129 10.478 \putrule from 5.129 10.478 to 5.105 10.478 \putrule from 5.105 10.478 to 5.091 10.478 \putrule from 5.091 10.478 to 5.084 10.478 \putrule from 5.084 10.478 to 5.080 10.478 }% \linethickness= 0.500pt \setplotsymbol ({\thinlinefont .}) {\color[rgb]{0,0,0}\plot 5.080 12.383 5.082 12.387 / \plot 5.082 12.387 5.086 12.393 / \plot 5.086 12.393 5.095 12.406 / \plot 5.095 12.406 5.110 12.427 / \plot 5.110 12.427 5.129 12.454 / \plot 5.129 12.454 5.154 12.493 / \plot 5.154 12.493 5.186 12.537 / \plot 5.186 12.537 5.222 12.588 / \plot 5.222 12.588 5.262 12.643 / \plot 5.262 12.643 5.306 12.704 / \plot 5.306 12.704 5.353 12.766 / \plot 5.353 12.766 5.402 12.829 / \plot 5.402 12.829 5.453 12.895 / \plot 5.453 12.895 5.503 12.958 / \plot 5.503 12.958 5.556 13.020 / \plot 5.556 13.020 5.609 13.079 / \plot 5.609 13.079 5.662 13.138 / \plot 5.662 13.138 5.715 13.193 / \plot 5.715 13.193 5.768 13.248 / \plot 5.768 13.248 5.823 13.299 / \plot 5.823 13.299 5.880 13.352 / \plot 5.880 13.352 5.937 13.401 / \plot 5.937 13.401 5.999 13.451 / \plot 5.999 13.451 6.062 13.500 / \plot 6.062 13.500 6.128 13.551 / \plot 6.128 13.551 6.198 13.602 / \plot 6.198 13.602 6.272 13.652 / \plot 6.272 13.652 6.322 13.686 / \plot 6.322 13.686 6.375 13.722 / \plot 6.375 13.722 6.430 13.758 / \plot 6.430 13.758 6.485 13.796 / \plot 6.485 13.796 6.545 13.832 / \plot 6.545 13.832 6.606 13.871 / \plot 6.606 13.871 6.672 13.911 / \plot 6.672 13.911 6.737 13.951 / \plot 6.737 13.951 6.805 13.991 / \plot 6.805 13.991 6.877 14.034 / \plot 6.877 14.034 6.949 14.074 / \plot 6.949 14.074 7.025 14.118 / \plot 7.025 14.118 7.104 14.161 / \plot 7.104 14.161 7.184 14.205 / \plot 7.184 14.205 7.269 14.252 / \plot 7.269 14.252 7.353 14.296 / \plot 7.353 14.296 7.440 14.343 / \plot 7.440 14.343 7.531 14.389 / \plot 7.531 14.389 7.622 14.436 / \plot 7.622 14.436 7.715 14.482 / \plot 7.715 14.482 7.808 14.529 / \plot 7.808 14.529 7.906 14.575 / \plot 7.906 14.575 8.003 14.624 / \plot 8.003 14.624 8.100 14.671 / \plot 8.100 14.671 8.200 14.717 / \plot 8.200 14.717 8.302 14.764 / \plot 8.302 14.764 8.401 14.810 / \plot 8.401 14.810 8.503 14.855 / \plot 8.503 14.855 8.604 14.899 / \plot 8.604 14.899 8.706 14.944 / \plot 8.706 14.944 8.807 14.988 / \plot 8.807 14.988 8.909 15.033 / \plot 8.909 15.033 9.011 15.075 / \plot 9.011 15.075 9.112 15.115 / \plot 9.112 15.115 9.214 15.157 / \plot 9.214 15.157 9.315 15.198 / \plot 9.315 15.198 9.415 15.236 / \plot 9.415 15.236 9.517 15.276 / \plot 9.517 15.276 9.618 15.314 / \plot 9.618 15.314 9.718 15.352 / \plot 9.718 15.352 9.819 15.388 / \plot 9.819 15.388 9.923 15.424 / \plot 9.923 15.424 10.020 15.460 / \plot 10.020 15.460 10.118 15.494 / \plot 10.118 15.494 10.219 15.528 / \plot 10.219 15.528 10.319 15.562 / \plot 10.319 15.562 10.420 15.596 / \plot 10.420 15.596 10.524 15.629 / \plot 10.524 15.629 10.628 15.663 / \plot 10.628 15.663 10.734 15.697 / \plot 10.734 15.697 10.842 15.731 / \plot 10.842 15.731 10.950 15.765 / \plot 10.950 15.765 11.060 15.797 / \plot 11.060 15.797 11.170 15.831 / \plot 11.170 15.831 11.280 15.864 / \plot 11.280 15.864 11.392 15.898 / \plot 11.392 15.898 11.504 15.932 / \plot 11.504 15.932 11.618 15.964 / \plot 11.618 15.964 11.731 15.998 / \plot 11.731 15.998 11.845 16.032 / \plot 11.845 16.032 11.957 16.063 / \plot 11.957 16.063 12.071 16.095 / \plot 12.071 16.095 12.184 16.127 / \plot 12.184 16.127 12.294 16.159 / \plot 12.294 16.159 12.406 16.190 / \plot 12.406 16.190 12.514 16.222 / \plot 12.514 16.222 12.622 16.252 / \plot 12.622 16.252 12.728 16.281 / \plot 12.728 16.281 12.831 16.311 / \plot 12.831 16.311 12.935 16.341 / \plot 12.935 16.341 13.034 16.368 / \plot 13.034 16.368 13.132 16.396 / \plot 13.132 16.396 13.227 16.423 / \plot 13.227 16.423 13.318 16.451 / \plot 13.318 16.451 13.409 16.476 / \plot 13.409 16.476 13.496 16.502 / \plot 13.496 16.502 13.578 16.525 / \plot 13.578 16.525 13.661 16.550 / \plot 13.661 16.550 13.739 16.574 / \plot 13.739 16.574 13.813 16.595 / \plot 13.813 16.595 13.887 16.618 / \plot 13.887 16.618 13.957 16.639 / \plot 13.957 16.639 14.023 16.660 / \plot 14.023 16.660 14.086 16.681 / \plot 14.086 16.681 14.150 16.703 / \plot 14.150 16.703 14.209 16.722 / \plot 14.209 16.722 14.311 16.758 / \plot 14.311 16.758 14.404 16.794 / \plot 14.404 16.794 14.491 16.828 / \plot 14.491 16.828 14.569 16.861 / \plot 14.569 16.861 14.641 16.895 / \plot 14.641 16.895 14.709 16.931 / \plot 14.709 16.931 14.770 16.967 / \plot 14.770 16.967 14.827 17.005 / \plot 14.827 17.005 14.880 17.043 / \plot 14.880 17.043 14.931 17.084 / \plot 14.931 17.084 14.978 17.124 / \plot 14.978 17.124 15.020 17.168 / \plot 15.020 17.168 15.060 17.211 / \plot 15.060 17.211 15.098 17.253 / \plot 15.098 17.253 15.130 17.295 / \plot 15.130 17.295 15.160 17.333 / \plot 15.160 17.333 15.183 17.369 / \plot 15.183 17.369 15.204 17.399 / \plot 15.204 17.399 15.219 17.424 / \plot 15.219 17.424 15.229 17.441 / \plot 15.229 17.441 15.236 17.454 / \plot 15.236 17.454 15.238 17.460 / \plot 15.238 17.460 15.240 17.462 / }% \put{\SetFigFont{6}{7.2}{\rmdefault}{\mddefault}{\updefault}{\color[rgb]{0,0,0}$\mathcal{W}-$}% } [lB] at 7.303 14.605 \put{\SetFigFont{6}{7.2}{\rmdefault}{\mddefault}{\updefault}{\color[rgb]{0,0,0}$L$}% } [lB] at 9.842 13.494 \put{\SetFigFont{6}{7.2}{\rmdefault}{\mddefault}{\updefault}{\color[rgb]{0,0,0}$\Gamma_d$}% } [lB] at 11.748 12.224 \put{\SetFigFont{6}{7.2}{\rmdefault}{\mddefault}{\updefault}{\color[rgb]{0,0,0}$\Gamma_u$}% } [lB] at 11.589 16.986 \put{\SetFigFont{6}{7.2}{\rmdefault}{\mddefault}{\updefault}{\color[rgb]{0,0,0}$\tilde{S}_-$}% } [lB] at 3.810 12.224 \put{\SetFigFont{6}{7.2}{\rmdefault}{\mddefault}{\updefault}{\color[rgb]{0,0,0}$\bar{N}_-$}% } [lB] at 3.810 10.319 \put{\SetFigFont{6}{7.2}{\rmdefault}{\mddefault}{\updefault}{\color[rgb]{0,0,0}$\tilde{N}_+$}% } [lB] at 15.399 17.304 \put{\SetFigFont{6}{7.2}{\rmdefault}{\mddefault}{\updefault}{\color[rgb]{0,0,0}$\tilde{S}_+$}% } [lB] at 15.399 15.399 \linethickness=0pt \putrectangle corners at 3.778 21.660 and 15.431 6.280 \endpicture} \end{center} \caption{Barrier for zero winding number} \end{figure} Thus $j(\xi)>0$ for all $\xi\in(-\frac{\pi}{2},\frac{\pi}{2})$. It follows that the slope of any orbit crossing the line is greater than the slope of the line. Thus orbits cannot cross this line from above to below. In particular, the orbit $\mathcal{W}^-$ starts at $\widetilde{S}_-$, which is above this line. Hence $\mathcal{W}^-$ cannot end at any copy of the node $N_+$ other than $\widetilde{N}_+$, so its winding number cannot be positive. Since the region ${\mathcal N}$ is connected, once $\mathcal{W}^-$ leaves ${\mathcal P}$ and enters ${\mathcal N}$, its $\Omega$ must decrease, hence $\mathcal{W}^-$ cannot end at any copy of the node $N_+$ that is higher than $\widetilde{N}_+$ either, and therefore the winding number of $\mathcal{W}^-$ cannot be negative, hence $w(\mathcal{W}^-) = 0$. \end{proof} Let $\lambda_0 = -1+a$. The above two propositions, in conjunction with the following immediate corollary of Proposition~\ref{prop:ssexists}, establish the existence a saddles connector ${\mathcal S}_0^\Omega(E_1,\lambda_0)$ forthe flow \refeq{dynsysOm}, for some $E_1\in(0,1)$: \begin{cor}\label{cor:om} Let $\lambda\leq -1+a$ be fixed. Suppose that there exists $0\leq E_1< E_2< 1$ such that the flow \refeq{dynsysOm} has corridors ${\mathcal K}_1(E_1,\lambda)$ and ${\mathcal K}_1(E_2,\lambda)$ with $w({\mathcal K}_1(E_1,\lambda))=0$ and $w({\mathcal K}_1(E_2,\lambda))\geq 1$. Then there is an $E\in(E_1,E_2)$ such that \refeq{dynsysOm} has a saddles connector ${\mathcal S}(E,\lambda)$. \end{cor} \begin{proof} Proposition~\ref{prop:ssexists} applies, with $E$ playing the role of the parameter $\mu$. \end{proof} Let $n\geq 1$. Suppose that given $\lambda_{n-1}\leq -1+a$ a saddles connector ${\mathcal S}^\Omega_{n-1}={\mathcal S}^\Omega(E_n,\lambda_{n-1})$ has been found for \refeq{dynsysOm}, for some $E_n\in(0,1)$. In the previous subsection we saw how this newly-found $E_n$ can be used to prove the existence of a saddles connector for the $\Theta$ flow \refeq{dynsysTh}, namely ${\mathcal S}^\Theta_n:= {\mathcal S}^\Theta_n(E_n,\lambda_n)$ for some $\lambda_n\leq -1+a$. Coming back to the $\Omega$ flow then, given the updated value $\lambda=\lambda_n$, a new saddles connector ${\mathcal S}_n^\Omega$ needs to be found with an updated energy $E_{n+1}$, {\em given that a saddles connector ${\mathcal S}^\Omega_{n-1}(E_{n},\lambda_{n-1})$ already exists.} Then, for all $\lambda'\in (\lambda,-1+a)$, there exists a corridor ${\mathcal K}_1(E,\lambda')$ of winding number $w({\mathcal K}_1) = 1$ for \refeq{dynsysOm}. More generally, we have \begin{thm}\label{thm:E} Fix $a\in(0,\frac{1}{2})$ and $\gamma\in (- \sqrt{2a(1-2a)},0)$. Then given any $\lambda\in[-1-a,-1+a]$, there exists a unique $$E = {\mathcal E}(\lambda)\in (0,1)$$ such that \refeq{dynsysOm} has a saddles connector ${\mathcal S}^\Omega(E,\lambda)$. Moreover, ${\mathcal E}$ is a $C^1$ function, and $|\frac{\partial {\mathcal E}}{\partial \lambda}| < \frac{1}{a}$. \end{thm} \begin{proof} Existence of a saddles connector is guaranteed by Propositions~\ref{prop:om1} and ~\ref{prop:om0}, and Corollary~\ref{cor:om}. To see uniqueness, suppose that there exists two saddles connectors ${\mathcal S}^\Omega(E,\lambda)$ and ${\mathcal S}^\Omega(E',\lambda)$ for \refeq{dynsysOm} for $E$ and $E'$ in $(0,1)$, and suppose $E<E'$. Let $\Omega_{E,\lambda}$ be the $\Omega$ component of ${\mathcal S}^\Omega(E,\lambda)$. We have $$ g_{E',\lambda}(\xi,\Omega_{E,\lambda}) = g_{E,\lambda}(\xi,\Omega_{E,\lambda})-2a(E'-E) <\dot{\Omega}_{E,\lambda}. $$ It thus follows that orbits of the $(E',\lambda)$ flow can only cross ${\mathcal S}^\Omega(E,\lambda)$ from above to below. On the other hand, since $E'>E$ the equilibrium point $\widetilde{S}_-(E',\lambda)$ is situated below $\widetilde{S}_-(E,\lambda)$, while $\widetilde{S}_+(E',\lambda)$ is above $\widetilde{S}_+(E,\lambda)$. Since ${\mathcal S}^\Omega(E',\lambda)$ coincides with both $\mathcal{W}^-(E',\lambda)$ and $\mathcal{W}^+(E',\lambda)$ it begins below ${\mathcal S}^\Omega(E,\lambda)$ and it ends above it, which is a contradiction, hence $E'=E$. Given $\lambda$, let ${\mathcal E}(\lambda)$ denote the unique value of $E$ for which a saddles connector ${\mathcal S}^\Omega({\mathcal E}(\lambda),\lambda) = (\xi(\tau),\Omega_{{\mathcal E}(\lambda),\lambda}(\tau))$ exists. We now prove that ${\mathcal E}$ is a $C^1$ function: Consider the two initial value problems for $\Omega^\pm(\tau)$, the $\Omega$ components of $\mathcal{W}^\pm$: $$ \dot{\Omega}^\pm = g_{E,\lambda}(\xi(\tau),\Omega^\pm),\qquad \Omega^-(-\infty) = -\pi+\cos^{-1}E,\quad \Omega^+(\infty) = -\cos^{-1}E $$ By standard ODE theory these two problems have unique smooth solutions $\Omega^\pm_{E,\lambda}(\tau)$ which also depend smoothly on the parameters $\lambda$ and $E$ (so long as $E<1$) for any finite $\tau$. Next recall that $\mathcal{W}^-$ is a saddles connector if it coincides with $\mathcal{W}^+$, which will be the case if these two orbits intersect at one point, e.g. if $\Omega_{E,\lambda}^+(0) = \Omega_{E,\lambda}^-(0)$. Let us define a smooth function \begin{equation}\label{eq:IFT} \Phi(E,\lambda) := \Omega^+_{E,\lambda}(0) -\Omega^-_{E,\lambda}(0) \end{equation} Let $\lambda_0\in[-1-a,-1+a]$ be fixed, and set $E_0={\mathcal E}(\lambda_0)$. Then $\Phi(E_0,\lambda_0) = 0$. By the Implicit Function Theorem, if \begin{equation}\label{cond:IFT} \frac{\partial \Phi}{\partial E} (E_0,\lambda_0) \ne 0, \end{equation} then there is a neighborhood ${\mathcal I}$ of $\lambda_0$ and a $C^1$ function $\tilde{{\mathcal E}}$ defined on ${\mathcal I}$ such that $\Phi(\tilde{{\mathcal E}}(\lambda),\lambda) =0$ for all $\lambda\in{\mathcal I}$. By the uniqueness result we have already shown, we must have $\tilde{{\mathcal E}} = {\mathcal E}$. Thus we only need to verify the condition \refeq{cond:IFT}. For $\lambda\in[-1-a,-1+a]$ and $E\in(0,1)$, let $$ u_\pm(\tau) := \frac{\partial}{\partial E} \Omega^\pm_{E,\lambda}(\tau). $$ Then $u_\pm$ satisfy the linear ODEs \begin{equation}\label{eq:v} \frac{du_\pm}{d\tau} = P_\pm(\tau) u_\pm -2a,\qquad P_\pm(\tau):=-2a\sin\xi(\tau)\sin\Omega^\pm_{E,\lambda}(\tau) + 2\lambda \cos\xi(\tau)\cos\Omega^\pm_{E,\lambda}(\tau) \end{equation} together with the initial conditions $$ u_-(-\infty) = \frac{-1}{\sqrt{1-E^2}},\qquad u_+(\infty) = \frac{1}{\sqrt{1-E^2}}. $$ Moreover, $$ \Phi(E,\lambda) = u_+(0) - u_-(0). $$ For $\tau_1^{},\tau_2^{}\in{\mathbb R}$ let $$ U_\pm(\tau_1^{},\tau_2^{}) := e^{-\int_{\tau_1^{}}^{\tau_2^{}} P_\pm(\tau) d\tau}. $$ Solving the ODEs \refeq{eq:v} for $u_\pm$ we obtain \begin{equation}\label{Us} U_\pm(\tau_1^{},\tau_2^{})u_\pm(\tau_2^{}) = u_\pm(\tau_1^{}) -2a \int_{\tau_1^{}}^{\tau_2^{}} U_\pm(\tau_1^{},\tau) d\tau. \end{equation} Note that $$ \lim_{\tau\to -\infty}P_-(\tau) = -2a\sqrt{1-E^2}< 0,\qquad \lim_{\tau\to \infty} P_+(\tau) = 2a\sqrt{1-E^2} > 0. $$ Thus for any fixed $\tau$, $$ U_-(\tau,\tau_1^{})\to 0 \mbox{ as } \tau_1^{}\to -\infty,\qquad U_+(\tau,\tau_2^{}) \to 0\mbox{ as } \tau_2^{} \to \infty. $$ And so, from \refeq{Us} we obtain \begin{equation}\label{solv} u_-(\tau) = -2a \int_{-\infty}^\tau U_-(\tau,\tau') d\tau',\qquad u_+(\tau) = 2a\int_\tau^\infty U_+(\tau,\tau') d\tau'. \end{equation} We know that $\Omega^+_{E_0,\lambda_0}(\tau) = \Omega^-_{E_0,\lambda_0}(\tau)$ for all $\tau$. Hence $U_+(\tau_1^{},\tau_2^{}) = U_- (\tau_1^{},\tau_2^{})=: U(\tau_1^{},\tau_2^{})$ and thus $$ \frac{\partial \Phi}{\partial E} (E_0,\lambda_0) = 2a\int_{-\infty}^\infty U(0,\tau') d\tau' >0 $$ so that \refeq{cond:IFT} is clearly satisfied. Since $\lambda_0$ was arbitrary we have shown that ${\mathcal E}\in C^1((-1-a,-1+a))$. We can furthermore compute the derivative of ${\mathcal E}$ by implicit differentiation. Let $$ v_\pm(\tau) := \frac{\partial}{\partial\lambda} \Omega^\pm_{E,\lambda}(\tau). $$ Then $v_\pm$ satisfy $$ \dot{v}_\pm = P_\pm(\tau) v_\pm + 2 \cos\xi(\tau)\sin\Omega^\pm_{E,\lambda}(\tau),\qquad v_-(-\infty) =0,\quad v_+(\infty) = 0. $$ Thus by a similar argument to above, $$ v_-(\tau) = \int_{-\infty}^\tau U_-(\tau,\tau') \cos\xi(\tau')\sin\Omega^-_{E,\lambda}(\tau') d\tau',\qquad v_+(\tau) = \int_\tau^\infty U_+(\tau,\tau') \cos\xi(\tau')\sin\Omega^+_{E,\lambda}(\tau') d\tau', $$ so that $$ \frac{\partial \Phi}{\partial \lambda}(E_0,\lambda_0) = -2\int_{-\infty}^\infty U(0,\tau') \cos\xi(\tau')\sin\Omega_{E_0,\lambda_0}(\tau') d\tau', $$ and thus $$ \frac{d{\mathcal E}}{d\lambda} = -\frac{{\partial \Phi}/{\partial \lambda}}{{\partial \Phi}/{\partial E}} = \frac{\int_{-\infty}^\infty U(0,\tau) \cos\xi(\tau)\sin\Omega_{E,\lambda}(\tau) d\tau}{a\int_{-\infty}^\infty U(0,\tau)d\tau}. $$ Moreover, clearly $$ \left|\int_{-\infty}^\infty U(0,\tau) \cos\xi(\tau)\sin\Omega_{E,\lambda}(\tau) d\tau\right| < \int_{-\infty}^\infty U(0,\tau)d\tau. $$ so that $$ \left|\frac{d{\mathcal E}}{d\lambda}\right| < \frac{1}{a}. $$ \end{proof} \subsubsection{The iteration argument} \begin{thm}\label{thereAREsaddlesaddleCONNECTORS} Let $a\in (0,\frac{1}{2})$ and $\gamma \in (- \sqrt{2a(1-2a)},0)$ be fixed. There exists a $\lambda\in [-1-a,-1+a] $ and $E \in (0,1)$ such that \refeq{dynsysTh} has a saddles connector ${\mathcal S}^\Theta(E,\lambda)$ and \eqref{dynsysOm} has a saddles connector ${\mathcal S}^\Omega(E,\lambda)$. \end{thm} \begin{proof} Set $\lambda_0 = -1+a$. For $n\geq 1$ let $$E_n := {\mathcal E}(\lambda_{n-1})\in(0,1),\qquad \lambda_n := \Lambda(E_n)\in [-1-a,-1+a]. $$ Thus $$ E_{n+1}= {\mathcal E}(\Lambda(E_{n})). $$ By Theorems~\ref{thm:E} and \ref{thm:La} we have $$ |E_{n+1} - E_n| \leq \delta |E_n- E_{n-1}| $$ where $$ \delta := \max_{0\leq E\leq1}\left|\frac{d\Lambda}{dE}\right| \max_{-1-a\leq \lambda\leq -1+a}\left| \frac{d{\mathcal E}}{d\lambda}\right| < a\cdot \frac{1}{a} = 1 $$ Thus by the contraction mapping theorem, the sequence $E_n$ converges, and thus so does the sequence $\lambda_n$. Let $\lambda = \lim_{n\to \infty} \lambda_n \in [-1-a,-1+a]$ and $E := \lim_{n\to \infty} E_n$. We must have $E = {\mathcal E}(\lambda)$ and thus $0<E<1$. \end{proof} \section{Summary and Outlook} We have studied the Dirac equation for a point electron in static, electromagnetic, flat spacetimes with Zipoy topology which include the zero-gravity limit of the electromagnetic Kerr--Newman spacetimes as special case, but which can feature a generalization of the Appell--Sommerfeld electromagnetic fields with any charge $Q$ and current $I$ one wants; the zero-$G$ Kerr--Newman spacetimes correspond to $Q=I\pi a$. In contrast to similar-spirited studies of the Dirac equation for a point electron on the Kerr--Newman spacetime, which are plagued by the presence of a Cauchy horizon and regions of closed timelike loops, \cite{KalMil1992,BelMar99,FinsterETalDperDNEerr,FinsterETalDperDNE,FinsterETalDKNa,FinsterETalDKNb,WINKLMEIERa,WINKLMEIERb,WINKLMEIERc,BelCac2010}, our zero-$G$ spacetimes do not possess any such physically troublesome features. Moreover, by working with the topologically non-trivial maximal analytical extension of z$G$KN and its electromagnetic fields, our treatment does not encounter physically troublesome problems like infinite charges, currents, and masses, which plague the topologically trivial Minkowski spacetime interpretations \cite{Isr70,Lyn04,GaiLyn07,Kai04}.\footnote{For $G>0$, the disk singularity of the ``single-leafed" truncation of the maximal analytically extended z$G$KN spacetime also features a negative mass density rotating at superluminal speed.} We proved that the spectrum of any self-adjoint extension of the pertinent Dirac Hamiltonian is symmetric about zero; this result holds for any charge $Q$ and current $I$ of the generalized z$G$KN spacetimes.\footnote{This result was to be expected, for on the z$G$KN spacetime the Dirac electron ``sees'' a charge $Q$ in one sheet and a charge $-Q$ in the other, and recalling that the Dirac operator for an electron in the Coulomb potential of a point proton in Minkowski spacetime has a symmetric continuous spectrum, plus a positive point spectrum which maps into a negative mirror image of it under the switch of the coupling constant $-e^2\to e^2$ (by charge conjugation).} We have also shown that the formal Dirac Hamiltonian on a spacelike slice of the maximal analytically extended static z$G$KN spacetime is essentially self-adjoint.\footnote{We note that by Stone's theorem there exists a unitary one-parameter group on the Hilbert space, generated by the unique self-adjoint extension of the Hamiltonian, which yields the time-evolution of the Dirac bi-spinors on the spacelike static slice of the z$G$KN spacetime. Thus the naked ring singularity does not cause any trouble.} We also showed that the self-adjoint Dirac operator on the z$G$KN spacetime has a continuous spectrum with a gap about zero which, under two smallness conditions, contains a pure point spectrum. Our results are far from exhaustive. In the following we list a number of interesting open problems which we hope will be solved in some future work. We begin with the Dirac point electron in z$G$KN spacetimes: \begin{itemize} \item{\textbf{Problem 1}}: Characterize the point spectrum of the Dirac Hamiltonian on z$G$KN in complete detail; to the extent possible, compute it analytically, or at least numerically in some representative situations. \begin{rem}\textit{ We suspect that there is a countably infinite set of energy eigenvalues which correspond with pairs of saddles connectors of arbitrary winding numbers $(w_1,w_2)\in{\mathbb Z}^2$; the possibility of such saddles connectors we already established, see Theorem \ref{thereAREsaddlesaddleCONNECTORS}. } \end{rem} \item {\textbf{Problem 2}}: Discuss the generalized scattering problem for Dirac spinor fields on the z$G$KN spacetimes. In particular, investigate the evolution when (part of) the Dirac spinor field ``dives'' through the ring from one sheet to the other. \end{itemize} We now come to the Dirac point electron in z$G$K spacetimes equipped with a generalization of the Appell--Sommerfeld electromagnetic fields to arbitrary charge $Q$ and current $I$: \begin{itemize} \item {\textbf{Problem 3}}: For the formal Dirac Hamiltonian on the $(Q,I)$-generalization of the z$G$KN spacetime (given $a$), show that essential self-adjointness holds if the ``coupling constant'' $(Q-I\pi a)e$ is small in magnitude; perhaps using a so-called Hardy--Dirac type estimate. \item{\textbf{Problem 4}}: Suppose essential self-adjointness fails if $|(Q-I\pi a)e|$ is too large. If so, what is the sharp constant for $|(Q-I\pi a)e|$? Determine the $(Q,I)$-parameter regimes (given $a$) in which the Dirac Hamiltonian on a generalization of the z$G$KN spacetime with Sommerfeld fields \refeq{def:AQI} has several self-adjoint extensions, respectively has no self-adjoint extension. Amongst the self-adjoint extensions, can one identify a distinguished one? \item{\textbf{Problem 5}}: In the cases of self-adjointness, characterize the spectrum of the Dirac Hamiltonian in complete detail; to the extent possible, compute the spectrum analytically, or at least numerically in some representative situations. In particular: \item{\textbf{Problem 5a}}: The continuous spectrum of quantum physical operator families is usually very robust. Show that Theorem \ref{thm:essspec} holds for the $(Q,I)$ generalization of our Dirac operators, at least as long as $(Q -I\pi a)e$ is small in magnitude. \item{\textbf{Problem 5b}}: Same consideration as above, for Theorem \ref{thm:ptspec}; thus: Can an eigenvalue of $\hat{H}$ on $\sf H$ for $Q =I\pi a$ be continuously deformed into an eigenvalue of $\hat{H}$ on $\sf H$ for $Q\ne I\pi a$ as long as $(Q -I\pi a)e$ is sufficiently small in magnitude? If so, does the size of the neighborhood of $Q=I\pi a$ into which an eigenvalue can be continued depend on the eigenvalue, or can one have a uniform control on the spectrum w.r.t. the coupling constant $(Q -I\pi a)e$? Is it possible that the point spectrum disappears completely if $(Q -I\pi a)e$ becomes too large in magnitude? \item {\textbf{Problem 6}}: Same as Problem 2, now for Dirac spinor fields on z$G$K equipped with generalizations of the Appell--Sommerfeld fields to arbitrary $Q$ and $I$. \end{itemize} So far our problems concern the Dirac equation on the zero-gravity limit case of the KN spacetimes, and its generalization to arbitrary $Q$ and $I$. To make contact with the existing studies of Dirac's electron in Kerr--Newman spacetimes, the following bifurcation problem suggests itself: \begin{itemize} \item {\textbf{Problem 7}}: Deform these zero-gravity spacetimes perturbatively by ``switching on'' $G$ and discuss the Dirac equation on them perturbatively as well. In particular, is it possible to perturb into generalizations of the Kerr--Newman spacetime with gyromagnetic ratios amounting to $g$-factors $g\neq g_{\mbox{\tiny{KN}}}^{}=2$, or is such a perturbation feasible only if $g =g_{\mbox{\tiny{KN}}}^{}=2$, viz. if $Q=I\pi a$? \end{itemize} In all the above problems, the energy(-density)-momentum(-density)-stress tensor is of classical electromagnetic nature. The Dirac spinor field does not influence the spacetime structure (in the zero-gravity limit the electromagnetic fields do not influence the spacetime structure, either). More to the point, Dirac's point electron is treated as a \emph{test particle} in this paper and in all the above problems, which should be a good approximation if the z$G$KN ring singularity is much more massive and can approximately be treated as ``infinitely massive.'' Yet, since test particles are only physical fiction, no matter how useful practically, an obvious task is to investigate the Dirac electron \emph{not} as a test particle. Thus: \begin{itemize} \item {\textbf{Problem 8}}: Treat the quantum-mechanical interaction of the Dirac electron with the ring singularity of the zero-$G$ Kerr--Newman spacetime symmetrically as a two-body problem. \begin{rem}\textit{ Of course, the same problem has not even been completely solved yet for the simpler setting of ``Dirac Hydrogen'' in Minkowski spacetime; i.e., how to go beyond the traditional textbook problem where the relativistic Hydrogen problem is treated by solving the Dirac equation of a point electron in flat Minkowski spacetime containing an \emph{infinitely massive} positive point charge (representing the proton). Traditionally the problem of finite mass of the proton (or nucleus, more generally) is addressed by perturbation theory, starting from Pauli's two-body equation and adding ``relativistic corrections'' in powers of $1/c$; or by perturbative QED-type calculations. The constrained relativistic two-body approach of Bethe--Salpeter \cite{BeSaBOOK} is perhaps the closest one has come to solving this problem. In a similar vein one may be able to take a finite ``ADM mass'' of the ring singularity into account. } \end{rem} \item {\textbf{Problem 9}}: Same as Problem 8, but now with the $(Q,I)$ generalizations of z$G$KN. \item {\textbf{Problem 10}}: Same as Problems 8 and 9, but now perturbatively for $G>0$. \item {\textbf{Problem 11}}: Compute the feedback of the Dirac spinor field onto the spacetime structure perturbatively when $G>0$. This amounts to the perturbative discussion of the so-called Einstein--Maxwell--Dirac system for small $G$. In this problem the energy-momentum-stress tensor, in addition to classical electromagnetic fields, also involves the Dirac spinor field which now influences the spacetime structure. \end{itemize} All these problems are difficult, and the amount of work needed to solve all of them can only be handled by the involvement of many mathematical physicists. In this vein we hope that our paper inspires some readers to join us in our pursuit. We ourselves have made some progress on problems 1, 3, 8, and 9, which we plan to report in forthcoming publications. \bigskip \noindent{\textbf{Acknowledgement:}} We thank Friedrich Hehl for his comments on an earlier version of this paper and for bringing several references to our attention. We are also thankful to the anonymous referee for constructive comments. \baselineskip=13pt \bibliographystyle{plain} \def$'${$'$}
\section{Introduction} Synapses connect neurons, allowing an action potential in the presynaptic neuron to influence the membrane potential of the postsynaptic neuron. The size of this membrane potential change is known as the postsynaptic potential (PSP), or alternatively the synaptic weight. Synaptic weights are updated based on information available locally at the synapse, for instance pre and postsynaptic activity. Theoretical neuroscientists have noted that there are three types of learning, supervised learning, reinforcement learning and unsupervised learning, corresponding to three essential cognitive tasks \cite{doya_complementary_2000,dayan_theoretical_2001}. First, supervised learning involves making predictions (e.g.\ about future events) from a set of inputs \cite{hastie_elements_2005}. For instance, a simple supervised learning task is to predict a single scalar output from a vector of inputs. To learn how to accomplish this task, the system repeatedly receives an input, makes a prediction and is given an error signal saying whether the prediction was too high or too low. It is believed that the cerebellum uses supervised learning to make short timescale predictions (e.g.\ predicting air puffs from aural cues \cite{kim_cerebellar_1997}) and to assist in fine motor control \cite{kandel_principles_2000}. Second, reinforcement learning involves learning behavioural responses that maximize reward (e.g.\ food reward) \cite{sutton_introduction_1998}. For instance, a simple reinforcement learning task is to map an input vector to an action represented by a single scalar value. To learn how to accomplish this task, the system repeatedly receives an input, chooses an action, and is told the reward for that action. It is believed that the basal ganglia uses reinforcement learning to perform the critical task of action selection \cite{dayan_theoretical_2001}. Finally, unsupervised learning involves finding structure in the inputs, perhaps by finding a direction in the input space with an interesting (non-Gaussian) distribution \cite{intrator_objective_1992}. It is believed that the cortex performs unsupervised learning, allowing us to understand and utilize the incredible volume of sensory data entering the brain \cite{koch_how_2006}. Local learning rules have been proposed for supervised \cite{dayan_theoretical_2001}, reinforcement \cite{sprekeler_code-specific_2009} and unsupervised \cite{hebb_organization_1949,rao_predictive_1999,zylberberg_sparse_2011} learning. These rules work well when given a large amount of data whose statistical properties do not change over time. However, being non-Bayesian, these rules do not exploit all locally available information in order to learn as rapidly as possible, which becomes problematic when there is less data available, or when the data changes over time. We therefore derived Bayesian synaptic learning rules. These learning rules do use all locally available information (up to approximations), are more effective than optimized classical learning rules, and make predictions about the results of plasticity experiments in active networks. \section{Results} We begin by writing down probabilistic generative models for the information received by a neuron. Then we derive Bayesian learning rules using the statistical structure given by the generative model. We demonstrate, in simulation, that our learning rules are far more effective than classical rules. Finally, our learning rules give strong predictions about the results of plasticity experiments under realistic patterns of network activity. \subsection{The model} The neuron's output, $y$, is a weighted sum of its imputs, $x_i$, plus noise, \begin{align} y &= \sum_i x_i w_i + \g_y \eta. \label{eq:gen:y} \end{align} where $\eta$ is standard Gaussian noise, $\P{\eta} = \N{\eta; 0, 1}$, and every future mention of $\eta$ is independently drawn standard Gaussian noise. In practice, PSPs, or equivalently synaptic weights, $w_i$, are noisy, in large part due to stochastic vesicle release \cite{branco_probability_2009}. We model these noisy PSPs using \begin{align} w_i &= m_i + \sqrt{k m_i} \eta_i, \end{align} where $m_i$ is the mean PSP and $k m_i$ is the variance. We chose this particular relationship between the mean and variance of PSPs in order to match experimental data \cite{song_highly_2005}. As $y$ is not constrained to be positive, we can think of $y$ as either the membrane potential, or the difference between the actual firing rate and some baseline firing rate. For supervised and reinforcement learning, the number of spikes, $x_i$, is drawn from a Poisson distribution, whose rates are chosen to mimic the distribution of rates in cortex (see Methods). These neurons receive a feedback signal, $f$, which is a deterministic function of a noisy error signal, \begin{align} \label{eq:def-delta} \delta = \yo - y + \g_f \eta. \intertext{Here, $\yo$ is the optimal output, which is defined in terms of the optimal weights, $\wosj$,} \label{eq:def-yo} \yo = \sum_j \wosj x_j. \end{align} Note that the neuron's eventual goal will be to set its average weights, $m_i$, as close as possible to the optimal weights, $\wosi$, so that the cell's outputs, $y$ are as close as possible to the optimal outputs, $\yo$. For supervised learning with continuous feedback, the feedback, $f$, is the noisy error signal, \begin{align} \label{eq:feedback-supervised-continuous} f = \delta. \end{align} For supervised learning with binary feedback, $f=1$ if the noisy error signal is above a threshold, $\theta$, $\theta$,and $f=-1$ otherwise, \begin{align} f| y, \yo &= \begin{cases} 1 & \text{if } \theta < \delta\\ -1 & \text{otherwise.} \end{cases} \label{eq:feedback-supervised-binary} \end{align} Finally, for reinforcement learning the feedback, or reward, reports the magnitude of the noisy error signal, but not its direction, \begin{align} f = -\abs{\delta}. \label{eq:feedback-policy} \end{align} In supervised and reinforcement learning, there was a feedback signal that depends on $\wo$, so the goal was to find $\wo$ from the feedback signal. In contrast, for unsupervised learning, there is no feedback signal. Instead, there is interesting statistical structure in the inputs, $\x$, that depends on $\wo$, so the goal is to find $\wo$ solely from the statistics of the inputs. For a Bayesian algorithm to find such interesting structure, (and to generate data on which to test our model neuron) we must first write down a generative model, describing how to generate inputs, $\x$ with the desired interesting structure. We chose to have the inputs, $\x$, be Gaussian in every direction except $\wo$. Along $\wo$, we set the distribution over $\x$ to be Laplacian. In order to enforce that structure, we generated $\yo$ on every time step from \begin{align} \label{P(yo)} \P{\yo} &= \text{Laplacian}\b{\yo; 0, b}, \end{align} where $b$ is chosen so that the variance along $\wo$ is similar to the variance in other directions. In order to generate a set of inputs, $\x$, that is consistent with both $\yo$ and the firing rate distribution, we sample $\x$ from, \begin{align} \label{P(x|yo)} \P{\x| \yo} &\propto \N{\x; \0, \L} \delta\b{\sum_j \wosj x_j - \yo} \intertext{where the diagonal elements of $\L$ are chosen to match the variance of a Poisson process with the appropriate rate, and the off-diagonal elements are set to $0$,} \L_{ij} &= \delta_{ij} \nu_i. \end{align} With these inputs any $x_i$ can be positive on one timestep and negative on another, so there is no distinction between excitatory and inhibitory cells. However, it is possible to interpret $x_i$ not as the inputs themselves, but as the difference between the actual input and the mean input. Using this interpretation, we could set the mean input to be large and positive for excitatory cells, and large and negative for inhibitory cells --- giving distinct excitatory and inhibitory populations. Finally, we do not expect the optimal weights to remain the same over all time, but to change slowly as the surrounding neural circuits and the statistics of the environment change. While these changes may actually be quite systematic, to a single synapse deep in the brain they will appear to be random. We formalise these random changes by assuming that the logarithm of the optimal weights, $\losi = \log \wosi$ follows a discretised Ornstein-Uhlenbeck process, \begin{align} \losi(t+1)| \losi(t) &= \b{1 - \frac{1}{\tau}} \b{\losi(t) - \mu_\prior} + \mu_\prior + \sqrt{\frac{2 \s_\prior^2}{\tau}} \n_i. \label{eq:dynamics} \end{align} where $t$, and the timescale $\tau$ are written in terms of time steps. We chose this particular noise process for three reasons. First, $\wosi = e^{\losi}$ takes only positive values, so excitatory weights cannot become inhibitory, and \textit{vice versa}. Second, spine sizes obey this stochastic process \cite{loewenstein_multiplicative_2011}, and while synaptic weights are not spine sizes, they are correlated \cite{matsuzaki_structural_2004}. Third, this noise process gives a log-normal stationary distribution of weights, as is observed \cite{song_highly_2005}. \subsection{Inference} The synapse's goal is to set its mean weight, $m_i(t+1)$, as close as possible to the optimal weight, $\wosi(t+1)$. Of course, the synapse does not know the optimal weight, so the best it can do is to set $m_i(t+1)$ equal to the expected value of the optimal weight given all past data, \begin{align} m_i(t+1) &= \E{\wosi(t+1)| \data_i(t)}. \end{align} Here, $\data_i(t) = \{d_i(t), d_i(t-1) \hdots d_i(1)\}$ is all past data, and $d_i(t)$ is the data at time step $t$. The data at time $t$, denoted $d_i(t)$, consists of the presynaptic activity, $x_i$, the postsynaptic activity, $y$, the actual PSP, $w_i$, a feedback signal, $f$ (if performing supervised or reinforcement learning), and information about the membrane time constant. Knowledge of the membrane time constant is not essential for learning, but proves to be useful for setting the learning rate on short timescales. In particular, the learning rate will turn out to depend on \begin{align} \label{sprior2} \slike^2 &= \sum_j x_j^2 \b{s_j^2 + k m_j}, \intertext{where} s_i^2 &= \Var{w_i| \data_i(t-1)}. \end{align} Notice that $\slike^2$, like the membrane time constant, is a measure of the overall level of input regardless of whether those inputs are excitatory or inhibitory, so, while $\slike^2$ and the membrane time constant are not exactly the same, the synapse should be able to estimate $\slike^2$ reasonably accurately from the membrane time constant, especially if the cell has many inputs. We therefore assume that $\slike^2$ is known by the synapse, even if, in reality, the cell has some uncertainty about the value of $\slike^2$. To compute $\P{\wosi(t+1)| \data_i(t)}$, and hence $\E{\wosi(t+1)| \data_i(t)}$, the neuron takes the distribution at the previous time step, $\P{\wosi(t)| \data_i(t-1)}$, uses Bayes theorem to incorporate new data, then takes into account the random changes in optimal weights across time. To take random changes into account, the neuron marginalises out the weights at the previous time step, \begin{align} \label{P(losi|D)-int} &\P{\wosi(t+1)| \data_i} = \int d \wosi \P{\wosi(t+1)| \wosi} \P{\wosi| \data_i}. \end{align} Here and in future, all quantities without an explicitly specified time are evaluated at time step $t$, so for instance $\wosi = \wosi(t)$. To compute $\P{\wosi| \data_i}$, the neuron must take its distribution at the previous time step, $\P{\wosi| \data_i(t-1)}$, and use Bayes theorem to incorporate the data received at timestep $t$, \begin{align} \label{P(losi|D)-bayes} \P{\wosi| \data_i} &\propto \P{d_i| \wosi} \P{\wosi| \data_i(t-1)}. \end{align} The resulting inferred distributions can be far too complex to deal with computationally, and certainly too complex for a synapse to work with, so we need to specify a family of approximate distributions. We chose to approximate the distribution over the log-weight, $\losi = \log \wosi$ using a Gaussian, with mean $\mu_i$, and variance $\s_i^2$, \begin{align} \label{P(losi|D)} \P{\losi| \data(t-1)} &= \N{\losi; \mu_i, \s_i^2}. \end{align} This approximation has two advantages. First, $\wosi = e^{\losi}$ takes only positive values, leading to learning rules that cannot, for instance, take an excitatory synapse and turn it inhibitory. Second, if the synapse is not given any data, then the dynamics (Equation~\eqref{eq:dynamics}) imply that the distribution over $\losi$ will be Gaussian, which can be captured perfectly by this approximating distribution. In order to set the parameters, $\mu_i$ and $\s_i^2$, of the approximate distribution, we do not need the full distribution, $\P{\losi(t+1)| \data_i}$, but only its mean and variance, which substantially simplifies the learning rules. In particular, for supervised and reinforcement learning the update rules are (see Methods for a full derivation) \begin{subequations} \label{eq:mu-sigma-update} \begin{align} \nonumber \mu_i(t+1) &= \E{\losi(t+1)| \data_i}\\ \label{eq:mu-update} &= \b{1 - \frac{1}{\tau}}\b{\mu_i + \b{\E{\delta| f, \slike^2, x_i, w_i} + x_i \b{w_i - m_i}} \frac{x_i m_i \s_i^2}{\slikei^2} - \mu_\prior} + \mu_\prior,\\ \nonumber \label{eq:sigma-update} \s_i^2(t+1) &= \Var{\losi(t+1)| \data_i}\\ &= \s_i^2 \b{1 - \frac{1}{\tau}}^2 \b{1 + \frac{\Var{\delta| f, \slike^2, x_i, w_i} - \slikei^2}{\slikei^2} \frac{\s_i^2 x_i^2 m_i^2}{\slikei^2}} + \frac{2 \s_\like^2}{\tau} \end{align} \end{subequations} \begin{align} \intertext{where} \label{eq:slikei2} \slikei^2 &= \slike^2 - x_i^2 k m_i, \end{align} and the exact form for $\E{\delta| f, \slike^2, x_i, w_i}$ and $\Var{\delta| f, \slike^2, x_i, w_i}$ are given in the methods. While these rules may look complex, the update rule for the mean is similar to the original delta rule \cite{dayan_theoretical_2001}. In particular, the change in weight is given by the product of three terms. First, $x_i$, ensures that the synpse only gets updated if the input was on. Second, there is an error signal, $\b{\E{\delta| f, \slike^2, x_i, w_i} + x_i \b{w_i - m_i}}$, stating whether this particular synaptic weight should increase or decrease. Third, there is an adaptive learning rate, $m_i \s_i^2/\slikei^2$. The learning rule for $\s_i^2$ is similar, though more difficult to interpret. \subsection{Simulations} Simulating each of these rules gives good results --- the inferred log-weight, $\mu_i$ tracks the optimal log-weight, $\losi$, with $\losi$ straying outside of the two standard deviation error bars about 5\% of the time (Figure~\ref{fig:time}). We used $1000$ inputs, and a time step of $10$~ms. For supervised learning we used a time constant, $\tau$, of $1,000$~s or around $15$ minutes, and for reinforcement and unsupervised learning we used a time constant of $10,000$~s or around $2 \tfrac{1}{2}$ hours. The firing rate distribution was chosen to be intermediate between relatively narrow ranges found by some \cite{oconnor_neural_2010}, and the extremely broad ranges found by others \cite{mizuseki_preconfigured_2013}. In particular, we use a log-normal distribution, with median at $1$~Hz, and with $95\%$ of firing rates being between $0.1$~Hz and $10$~Hz. The parameters of the dynamics, $\mu_\prior$ and $\s_\prior^2$ were chosen by fitting a lognormal distribution to measured synaptic weights \cite{song_highly_2005}. \begin{figure} \centering \includegraphics[width=0.5\textwidth]{time.pdf} \caption{Bayesian learning rules track the true weight, and estimate uncertainty. The blue line is the true weight, the red line represents the mean of the inferred distribution, $e^{\mu_i}$, and the red area represents two standard deviation error bars, $e^{\mu_i \pm 2 \s_i}$. The time course is $3$ time constants. \textbf{A} Supervised learning, continuous feedback. \textbf{B} Supervised learning, binary feedback. \textbf{C} Reinforcement learning. \textbf{D} Unsupervised learning. } \label{fig:time} \end{figure} We found that our Bayesian learning rules are more effective than standard classical learning rules with any learning rate (Figure~\ref{fig:comparison}). We evaluated performance using the mean square error (MSE) between $\wosi$ and $m_i$. More details, including the classical learning rules used for each type of learning are given in the Methods. \begin{figure} \centering \includegraphics[width=0.5\textwidth]{comparison.pdf} \caption{ Bayesian learning rules have a lower mean square error than classical learning rules. The black line is the mean square error for the classical learning rule with a given learning rate (x-axis), divided by the mean square error for the Bayesian learning rule, which is constant as it is parameter-free. The red line is 1. \textbf{A} Supervised learning, continuous feedback. \textbf{B} Supervised learning, binary feedback. \textbf{C} Reinforcement learning. \textbf{D} Unsupervised learning. } \label{fig:comparison} \end{figure} \subsection{Predictions} Our learning rules make two predictions, both of which are intuitively sensible and experimentally testable. First, the learning rate should vary across time, falling when many inputs are active, and rising when few inputs are active (in our equations, this occurs because of the $1/\slikei$ term in Equation~\eqref{eq:mu-update}). We can see that this prediction is intuitively sensible by considering a time step upon which only one input fires. In this case, the synaptic weight associated with the input that fired is entirely responsible for any error, so that synaptic weight should be updated drastically. In contrast, if many cells fire then it is not clear which synaptic weights are responsible for an error, so it is sensible to make only small updates to the weights. In our learning rules this occurs because $\slike^2$, which appears in the denominator of the learning rate in Equation~\eqref{eq:mu-update}, increases as the number of active inputs increases, (Equation~\eqref{sprior2}). We therefore predict fast learning during plasticity protocols in silent networks, when only the stimulated cells are active, and slow learning in active networks. There is some circumstantial evidence in favour of this prediction, in that weights can be changed within a few minutes during plasticity experiments in silent slices \cite{sjostrom_rate_2001}, whereas weight changes \textit{in vivo}, in active networks occur over much longer timescales \cite{loewenstein_multiplicative_2011}. However, a direct test of this prediction remains necessary. Second, learning rates should vary across synapses, being lower for synapses whose pre-synaptic cells have a higher average firing rate. We can see that this is intuitively sensible by comparing a synapse whose presynaptic cell fires rarely to a synapse whose presynaptic cell fires frequently. If the pre-synaptic cell fires rarely, the synapse is given only a few opportunities to update the weight, so should make large updates. In contrast, if the presynaptic cell fires rapidly, then the synapse is given many opportunities to update the weight, so should update the weight only a little for each presynaptic spike. In our learning rules, this occurs because lower firing rates give the cell fewer opportunities to learn, leading to higher uncertainty and hence higher learning rates. More formally, we show in the Methods that, \begin{align} \label{eq:ltp} \frac{\delta m_i}{m_i} \propto \frac{s_i^2}{m_i} \propto \frac{1}{\sqrt{\nu_i}}. \end{align} The first relationship comes from manipulating the learning rules for the mean (Equation~\eqref{eq:mu-update}), and the second comes from mean field calculations detailed in Methods, and confimed in simultations (Figure~\ref{fig:synapses}). \begin{figure} \centering \includegraphics[width=0.5\textwidth]{synapses.pdf} \caption{Uncertainty, and hence learning rate, increases as firing rate decreases. \textbf{A} Supervised learning, continuous feedback. \textbf{B} Supervised learning, binary feedback. \textbf{C} Reinforcement learning. \textbf{D} Unsupervised learning. } \label{fig:synapses} \end{figure} \section{Discussion} We derived Bayes optimal learning rules, which allow a synpase to exploit all locally available information in order to learn as rapidly as possible. These rules allow synapses to learn far more rapidly than optimized classical methods when performing supervised, unsupervised and reinforcement learning tasks. Furthermore, our Bayes optimal learning rules suggest that learning rates should vary across time and across synapses. In particular, learning rates should be higher at times when there are active fewer inputs into a cell, and learning rates should be higher for synapses whose presynaptic cells fire more infrequently. Both of these predictions can, in principle, be tested, something that we hope will occur in the next generation of plasticity experiments. Out of necessity, we have made a variety of specific choices about the information synapses have available to them, and about the statistical properties of a neuron's inputs. However, our framework is unique in providing the machinery to derive a learning rule for any combination of locally available information and input statistics. There are therefore two important directions for future work: considering more complex and realistic locally available information, and considering more complex and realistic input statistics. There are many directions to explore concerning the information available to a synapse. First, the synapse might have access to more information than we assumed, perhaps it might have some information about the activity of the surrounding synapses, allowing the synapse to learn something about the correlations between its weight and surrounding weights. Second, the synapse might have access to less information than we assumed, perhaps the postsynaptic activity does not backpropagate effectively to the synapse, or the synapse is not aware of a presynaptic spike because of a vesicle release failure. Finally, with a detailed characterisation of the electrical properties of dendrites, it may be possible to find a learning rule that is based only on the membrane potential and calcium signals at the synapse. In each of these cases the resulting learning rules may well make interesting and testable predictions. Our choices about the statistics of the inputs were intended to be a first step, so there are again many interesting directions to explore. Neural networks exhibit a variety of different global behaviours, from oscillations with changing power and frequency \cite{ray_differences_2010} to up-down states \cite{destexhe_are_2007}. Furthermore, there are debates about the structure of neural activity --- whether it is correlated \cite{ecker_decorrelated_2010} or whether it exists on a low-dimensional manifold \cite{byron_gaussian-process_2009}. When considering learning in realistic neural networks, it will be important for neurons and synapses to take into account both the structure and state of neural activity. The flexibility of our approach provides a means to connect relatively abstract theory to complex biological reality, an approach that, we hope, will significantly increase our understanding of synaptic plasticity. \section*{Acknowledgements} We would like to acknowledge Jesper Sj\"ostrum for valuable feedback on this manuscript. This work was supported by the Gatsby Charitable Foundation.
\section{Introduction} Even small amount of disorder can change considerably some properties of condensed matter systems. The most famous examples are probably the Anderson localization \cite{Evers} and the Kondo effect \cite{Hewson}. Disordered boson systems (so-called dirty-boson systems) have attracted much attention recently because a possibility of studying some peculiar predictions in this field has arisen in magnetically disordered spin-liquid-like materials and in optical lattices of ultracold atoms (see Ref.~\cite{Zhel13} for review). In particular, the existence of a disordered gapless Bose-glass (BG) phase was predicted for dirty bosons between gapped Mott-insulating (MI) and gapless superfluid (SF) phases. \cite{Fisher} A general theorem has been proven recently which states that BG phase always intervenes between MI and SF phases. \cite{theorem} The transition between MI and BG phases takes place via the Griffiths mechanism. \cite{grif} The nature of the quantum phase transition from BG to SF phases has been widely debated in recent years (see Refs.~\cite{Zhel13,kisel} and references therein). It has been understood recently that spin-1 magnets with large single-ion easy-plane anisotropy $\cal D$ and spin-$\frac12$ dimerized systems are convenient objects for discussing the dirty boson problem if disorder is realized in exchange coupling constants and/or $\cal D$. \cite{Zhel13} Such systems can be prepared in practice by creating a disorder on peripheral sites involved in superexchange interactions. A number of both large-$\cal D$ and dimerized substances with such disorder have been synthesized to date. \cite{Zhel13} At small magnetic field $H$, pure systems of this type have singlet ground states separated from the triplet excitation bands by gaps. For quasi-1D, quasi-2D and 3D systems, the phase diagram on the $T-H$ plane is presented in Fig.~\ref{th}(a) that shows a magnetically ordered gapless (SF) phase at $H_{c1}<H<H_{c2}$ and paramagnetic gapped (MI) phases at $H<H_{c1}$ and $H>H_{c2}$ (the fully polarized phase). \begin{figure} \noindent \includegraphics[scale=0.8]{TH-diagramSD.eps} \includegraphics[scale=0.8]{TH-BG1.eps} \hfil \caption{Phase diagrams of (quasi-)3D dimerized spin-$\frac12$ systems and spin-1 magnets with large single-ion easy-plane anisotropy $\cal D$ in magnetic field $H$. (a) Systems without defects with a canted magnetic ordering inside the dome. (b) Systems with a small fraction of bonds with strengthen and/or weaken exchange coupling constants (and/or $\cal D$ value). Bose-glass phases are denoted as BG.} \label{th} \end{figure} To explain the problem we address in the present paper, let us strengthen some randomly chosen intradimer exchange coupling constants $\cal J$ (or $\cal D$ values on some sites) in these systems. Localized impurity levels can appear inside the gap in the fully polarized phase which start to "condense" at some critical field $H_{bg2}>H_{c2}$ transferring the system into the BG phase (see Fig.~\ref{th}(b)). There are magnetically ordered islands around "strong" defects in this high-field BG phase which are well separated from each other by a nonmagnetic background, while no coherent long-range magnetic order exists in the whole system. Local quantized axes align in all islands (or all islands merge) when the transition to the magnetically ordered phase takes place at $H=H_{c2}$. \cite{foot} In contrast to the fully polarized phase, there are no localized impurity levels inside the gap for "strong" defects at small $H$. Nevertheless the general theorem \cite{theorem} requires that the field-induced transition to the ordered phase should take place via a BG phase. On the other hand, the Zeeman term commutes with the Hamiltonian and the magnetic field plays a role of a chemical potential at small and large $H$ in bosonic analogues of spin Hamiltonians (see also below for detail). Then, one is lead to a somewhat counterintuitive conclusion that at least low-energy states in the excitation band are localized at $H=0$ in the case of "strong" defects and their "condensation" drives the system into the BG phase at $H_{bg1}<H_{c1}$ (see Fig.~\ref{th}(b)). \cite{foot} Analogously, one leads to the same counterintuitive conclusion for "weak" defects at large $H$. As a result a natural question arises: which of the states in the band become localized and which of them remain propagating acquiring only a finite damping due to scattering on defects. This question looks particularly important in the light of recent excitation spectra measurements in IPA-Cu(Cl$_x$Br$_{1-x}$)$_3$ (Ref.~\cite{nafradi}) and $\rm (C_4H_{12}N_2)Cu_2(Cl$$_{1-x}$Br$_x$)$_6$ (Ref.~\cite{huv2}), dimerized materials with no impurity levels inside the gap at $H=0$. Despite considerable interest to bond disorder in spin-liquid-like magnets, this question has not been raised yet \cite{Zhel13,Vojta,Arlego} and the possibility of localization of some states in the band has not been considered in the experimental papers Refs.~\cite{nafradi,huv2}. We attack this problem analytically using the conventional $T$-matrix approach that is widely used in discussion of defects in condensed matter theory \cite{donia} and was proven to be very useful for magnetically ordered systems with impurities. \cite{Izyumov,white,orderdif,wan,2dvac,referee1,referee2,igar} This approach allows to find corrections to Green's functions, the excitation spectrum and the density of states (DOS) in the first order in defects concentration $c\ll1$. If the expansion in terms of $c$ is valid, excitations remain propagating in disordered systems although a finite damping arises due to scattering on defects. It can happen, however, that terms of higher-order in $c$ are also important for some momenta $\bf k$ signifying analytical approach inapplicability and necessity of an additional analysis. This can be a sign of propagating modes resonance scattering on defects (which, however, can remain propagating as a result of this scattering) \cite{Izyumov,white} or a localization of some states (see, e.g., Ref.~\cite{2dvac}). In dimerized spin-$\frac12$ systems and in integer-spin large-$\cal D$ magnets, the $T$-matrix approach allows to perform a unified consideration of all gapped phases, because Green's functions and spectra of propagating excitations have the same form. The only formal requirement to be fulfilled is that excitations in pure systems are weakly interacting. This condition holds at small $H$ if the intradimer exchange constant $\cal J$ and $\cal D$ are much larger than other exchange coupling constants $J_{ij}$. In the fully polarized phase, the magnon interaction does not lead to renormalization of observables at $T=0$ and it can be omitted. \cite{chub} We consider below the disorder in $\cal J$ or $\cal D$ as well as in $J_{ij}$. It is found that the analytical approach is invalid in 1D, 2D and 3D systems for states near the bottom and the top of excitation bands for all kinds of bond disorder (i.e., for "strong" and "weak" defects and for systems containing both "strong" and "weak" impurities). The linear size of regions in the $\bf k$-space inside which the analytical approach does not work scale as some powers of $c$. To clarify the nature of states near band edges, we perform numerical calculations for 1D and 2D systems which show that these states are {\it localized} and they have nothing to do with conventional wavepackets. In 2D systems, the numerical analysis shows that states inside the band far from its edges are well-defined wavepackets which energies and lifetimes are given by analytical expressions obtained in the first order in $c$ (one expects the same conclusion in 3D systems). In 1D systems, {\it all} states in the band are found to be localized (similar to 1D electronic systems). At the same time, some of the states inside the band reflect properties of propagating short-wavelength excitations which energies and lifetimes are given by analytical expressions obtained in the first order in $c$. Besides, it is found that some states inside the band in 1D systems are not conventional wavepackets due to a resonant scattering on strong enough defects. Our spectrum calculations show that in the vicinity of $H_{bg1}$ or $H_{bg2}$, if no localized impurity levels exist in the gap, the ratio of the long-wavelength propagating modes damping $\gamma_{\bf k}$ to their energy $\varepsilon_{\bf k}$ can reach $c/k^2$ in the range of this result validity $1\gg k\gg \sqrt c$. This contrasts with magnetically ordered gapless magnets in which $\gamma_{\bf k}/\varepsilon_{\bf k}$ does not exceed $c$. \cite{wan,2dvac,syromyat1,syromyat2} Thus, the damping of propagating excitations can be much more pronounced in considered systems than in magnetically ordered gapless magnets with impurities. The results obtained can be relevant to other gapped phases in bond disordered spin systems both with and without a long range magnetic order (e.g., bond disordered easy axis ferromagnets and antiferromagnets with large easy axis anisotropy). Our main analytical results are represented in quite a model-independent form that allows using them in analysis of other systems. The rest of the present paper is organized as follows. Pure systems are considered in Sec.~\ref{puresec}, where we derive bosonic analogs of spin Hamiltonians in all gapped phases using standard spin operators representations and demonstrate their similarity in dimerized and large-$\cal D$ systems. Our analytical and numerical methods are described in Sec.~\ref{method}. Bond disordered systems are considered in Sec.~\ref{disordersec}. Sec.~\ref{sum} contains a summary of results and our conclusions. An appendix is added with details of calculations. \section{Gapped phases in pure systems} \label{puresec} In this section, we derive Bose-analogs of spin Hamiltonians describing dimerized and large-$\cal D$ systems at $H<H_{c1}$ and $H>H_{c2}$ and demonstrate their similarity that allows the subsequent unified consideration. We derive elementary excitation spectra neglecting interaction between quasiparticles. This harmonic approximation is justified at $H>H_{c2}$ because spin-wave interaction does not modify one-particle Green's functions in the fully polarized phase. \cite{chub} At $H<H_{c1}$, the quasiparticle interaction can be neglected in the first order in the small exchange coupling $J_{ij}$ of spins from different dimers or from different sites (in large-$\cal D$ systems). It is shown below that spectra of all modes in this approximation have the form \begin{equation} \varepsilon_{\bf k} = \Delta + \frac a2 (J_{\bf k}-J_{\bf k_0}), \label{spec} \end{equation} where $a>0$ is a constant, $\Delta$ is the gap value, $J_\mathbf{k}$ is the Fourier transform of $J_{ij}$, and ${\bf k}_0$ is the momentum at which $J_{\bf k}$ reaches its minimum. For simplicity, we assume below that $J_{ij}\ne0$ for nearest neighbors only and that $J_{ij}$ either positive or negative so that all components of ${\bf k}_0$ are equal to $\pi$ if $J_{ij}>0$ and ${\bf k}_0={\bf 0}$ when $J_{ij}<0$. Then, $\varepsilon_{\bf k}$ depends quadratically on $\kap={\bf k}-{\bf k}_0$ near its minimum: \begin{equation} \varepsilon_{\bf k} = \Delta + \frac a2|J| \kappa^2. \label{spec0} \end{equation} One obtains similar quadratic dependence near the spectrum maximum: $\varepsilon_{\bf k} = \Delta +a|J_{\bf k_0}|- a|J| \kappa^2/2$. Here and below $\kap$ measures a deviation of the momentum from values at which the bare spectrum has a minimum or a maximum. All the results obtained in this section are not original. We omit some details of the corresponding consideration which can be found in cited papers. \subsection{Spin-1/2 dimer systems} We discuss 1D, 2D and 3D simple Bravias lattices of spin-$\frac12$ dimers which Hamiltonian is written in the following form: \begin{equation} \mathcal{H}=\sum_i {\cal J} {\mathbf S}_{i,1} \cdot \mathbf{S}_{i,2} + \sum_{ m} \sum_{\langle i,j\rangle} J_{ij} \left( \mathbf{S}_{i,1} \cdot \mathbf{S}_{j,1} + \mathbf{S}_{i,2} \cdot \mathbf{S}_{j,2}\right) - h \sum_i \left(S^z_{i,1}+S^z_{i,2} \right), \label{ham1} \end{equation} where $\mathbf{S}_{i,n}$ denotes $n$-th spin ($n=1,2$) in $i$-th dimer, $h=g\mu_BH$ is the external magnetic field and $\langle i,j\rangle$ denote nearest neighbor dimers. We set below the intradimer coupling constant ${\cal J}=1$. The exchange coupling between spins from different dimers in Eq.~\eqref{ham1} is taken in the simplest form. \subsubsection{$H<H_{c1}$} The system has a singlet ground state that corresponds to the paramagnetic phase in Fig.~\ref{th}(a). We derive the Bose-analog of spin Hamiltonian \eqref{ham1} in the standard way \cite{sach} by introducing three Bose-operators $\mathfrak a$, $\mathfrak b$, and $\mathfrak c$ for each dimerized bond which act on the vacuum spin state $|0\rangle=\frac{1}{\sqrt{2}}\left(|\uparrow\downarrow\rangle-|\downarrow\uparrow\rangle \right)$ as follows: $\mathfrak a|0\rangle=\mathfrak b|0\rangle=\mathfrak c|0\rangle=0$, $\mathfrak a^+|0\rangle=|\uparrow\uparrow\rangle$, $\mathfrak b^+|0\rangle=|\downarrow\downarrow\rangle$, and $\mathfrak c^+|0\rangle=\frac{1}{\sqrt{2}}\left(|\uparrow\downarrow\rangle+|\downarrow\uparrow\rangle \right)$. One has for spin operators \begin{equation} \begin{split} S^+_{i,1}=\frac{1}{\sqrt{2}}(\mathfrak a_i^+(\mathfrak c_i-1)+(\mathfrak c_i^++1)\mathfrak b_i),\\ S^+_{i,2}=\frac{1}{\sqrt{2}}(\mathfrak a_i^+(\mathfrak c_i+1)+(\mathfrak c_i^+-1)\mathfrak b_i),\\ S^z_{i,1}=\frac{1}{2}((\mathfrak c_i^++\mathfrak c_i)+\mathfrak a_i^+\mathfrak a_i-\mathfrak b_i^+\mathfrak b_i),\\ S^z_{i,2}=\frac{1}{2}(-(\mathfrak c_i^++\mathfrak c_i)+\mathfrak a_i^+\mathfrak a_i-\mathfrak b_i^+\mathfrak b_i). \end{split} \label{spinrep} \end{equation} To fulfill the requirement that no more than one triplon $\mathfrak a$, $\mathfrak b$ or $\mathfrak c$ can sit on the same bond, one has to introduce constraint terms into the Hamiltonian which describe an infinite repulsion between triplons $U \sum_i(\mathfrak a^+_i \mathfrak a^+_i \mathfrak a_i \mathfrak a_i + \mathfrak b^+_i \mathfrak b^+_i \mathfrak b_i \mathfrak b_i + \mathfrak c^+_i \mathfrak c^+_i \mathfrak c_i \mathfrak c_i + \mathfrak a^+_i \mathfrak b^+_i \mathfrak a_i \mathfrak b_i + \mathfrak a^+_i \mathfrak c^+_i \mathfrak a_i \mathfrak c_i + \mathfrak b^+_i \mathfrak c^+_i \mathfrak b_i \mathfrak c_i)$, where $U \to +\infty$. After substituting Eqs.~\eqref{spinrep} into Eq.~\eqref{ham1} one obtains the Bose-analog of the spin Hamiltonian which contains the constant term and terms with products of two and four Bose-operators. We restrict ourselves below by calculating triplon spectra in the first order in the interdimer coupling $J_{ij}$. It can be shown (see, e.g., Ref.~\cite{oleg}) that triplons spectra are defined only by bilinear part of the Hamiltonian in the first order in $J_{ij}$ and one has to take into account quasiparticles interaction to find spectra in higher orders. Then, the bilinear part of the Hamiltonian \begin{equation} \mathcal{H}_2 = \sum_{{\bf k}} \left[ \left( 1 + \frac{J_{\bf k}}{2}-h\right)\mathfrak a^+_{\bf k} \mathfrak a_{\bf k} + \left( 1 + \frac{J_{\bf k}}{2}+h\right)\mathfrak b^+_{\bf k} \mathfrak b_{\bf k} + \left( 1 + \frac{J_{\bf k}}{2}\right)\mathfrak c^+_{\bf k} \mathfrak c_{\bf k} - \frac{J_{\bf k}}{2} (\mathfrak a_{\bf k} \mathfrak b_{-{\bf k}} + \mathfrak a^+_{\bf k} \mathfrak b^+_{-{\bf k}}) +\frac{J_{\bf k}}{4}(\mathfrak c_{\bf k} \mathfrak c_{-{\bf k}} + \mathfrak c^+_{\bf k} \mathfrak c^+_{-{\bf k}}) \right] \label{h2} \end{equation} gives Eq.~\eqref{spec} for triplons spectra in the first order in $J_{ij}$ with $a=1$, $\Delta_{\mathfrak a} = 1-h+\frac12J_{{\bf k}_0}$, $\Delta_{\mathfrak b} = 1+h+\frac12J_{{\bf k}_0}$, and $\Delta_{\mathfrak c} = 1+\frac12J_{{\bf k}_0}$. As the last two terms in Eq.~\eqref{h2} do not contribute to spectra in the first order in $J_{ij}$, we omit them in the subsequent consideration. \subsubsection{$H>H_{c2}$} One can use the Holstein-Primakoff spin representation in the fully polarized phase at $H>H_{c2}$. As soon as magnon interaction does not lead to spectrum renormalization in this case, \cite{chub} we restrict ourselves with the linear spin-wave approximation and use the following expressions: \begin{equation} S^x_{i,n}=\frac{1}{2}(\mathfrak a_{i,n}+\mathfrak a^+_{i,n}), \quad S^y_{i,n}=-\frac{i}{2}(\mathfrak a_{i,n}-\mathfrak a^+_{i,n}), \quad S^z_{i,n}=\frac{1}{2}-\mathfrak a^+_{i,n}\mathfrak a_{i,n}. \label{hp} \end{equation} After the Hamiltonian transformation and introduction of new Bose-operators \begin{equation} \mathfrak a_{i,I}=\frac{\mathfrak a_{i,1}+\mathfrak a_{i,2}}{\sqrt{2}}\quad \mbox{and}\quad \mathfrak a_{i,II}=\frac{\mathfrak a_{i,1}-\mathfrak a_{i,2}}{\sqrt{2}}, \end{equation} one obtains \begin{equation} \mathcal{H}_2 = \sum_{{\bf k}} \left[ \left(h-\frac12J_{\bf 0}+\frac12J_{\bf k}\right)\mathfrak a^+_{{\bf k},I}\mathfrak a_{{\bf k},I}^{} + \left(h-1-\frac12J_{\bf 0}+\frac12J_{\bf k}\right)\mathfrak a^+_{{\bf k},II}\mathfrak a_{{\bf k},II}^{}\right], \label{h22} \end{equation} where two branches of elementary excitations have spectra of the form \eqref{spec} with $a=1$. \subsection{Systems with integer spin and large single-ion easy-plane anisotropy} We consider the following Hamiltonian for such systems \begin{equation} \mathcal{H}=\sum_{\langle i,j\rangle} J_{ij} \mathbf{S}_i \cdot \mathbf{S}_j + {\cal D} \sum_{i} (S^z_i)^2 - h \sum_i S^z_{i}, \label{ham2} \end{equation} where ${\cal D}>0$ and ${\cal D} \gg |J_{ij}|$. Similar to spin-dimer systems, these ones have singlet (paramagnetic) ground states at small $h$ in which all spins are predominantly in states with $S^z=0$. If $S=1$, the system has the $T$--$H$ diagram shown in Fig.~\ref{th}(a). For greater integer $S$, the phase diagram contains $S$ separated (if $|J_{ij}|$ is small enough) regions with canted magnetic ordering. \cite{chub} In this case, notations $H_{c1}$ and $H_{c2}$ used below denote the smallest and the largest critical fields, respectively. Similar to the intradimer coupling constant $\cal J$, we set below ${\cal D}=1$. \subsubsection{$H<H_{c1}$} For arbitrary integer spin $S$, the Bose-analog of spin Hamiltonian \eqref{ham2} can be derived using the following representation (see Ref.~\cite{Sizanov} for details): \begin{eqnarray} S^+_i &=& \mathfrak b^+_i (f_1-f_2 \mathfrak b^+_i \mathfrak b_i)+(f_1-f_2 \mathfrak a^+_i \mathfrak a_i)\mathfrak a_i,\nonumber\\ S^z_i &=& \mathfrak b^+_i \mathfrak b_i -\mathfrak a^+_i \mathfrak a_i, \label{drep} \end{eqnarray} where two types of Bose-operators are introduced and \begin{equation} f_1=\sqrt{S(S+1)}, \quad f_2=\sqrt{S(S+1)}-\sqrt{(S-1)(S+2)/2}>0. \end{equation} To obtain quasiparticles spectra in the first order in the exchange interaction, one needs only the bilinear part of the Hamiltonian \cite{Sizanov} which has the form \begin{equation} \mathcal{H}_2 = \sum_{{\bf k}} \left[ \left(1 + h + \frac{f^2_1}{2}J_{\bf k}\right)\mathfrak a^+_{\bf k} \mathfrak a_{\bf k} + \left(1 - h + \frac{f^2_1}{2}J_{\bf k}\right)\mathfrak b^+_{\bf k} \mathfrak b_{\bf k}\right] \label{h23} \end{equation} and which describes two branches of excitations with spectra \eqref{spec}, where $a=f^2_1=S(S+1)$. \subsubsection{$H>H_{c2}$} The high-field fully polarized phase can be considered using the Holstein-Primakoff spin representation that gives for the Hamiltonian in the linear spin-wave approximation \begin{equation} \mathcal{H}_2 = \sum_{{\bf k}} \left(h-(2S-1)-S(J_{\bf 0}-J_{\bf k})\right)\mathfrak a^+_{\bf k} \mathfrak a_{\bf k}. \label{h24} \end{equation} The spectrum has the form \eqref{spec} in this case with $a=2S$. It should be noted that magnetic field can only change the gap value but does not affect the strength of the quasiparticles interaction at $T=0$ and plays the role of a chemical potential. This is related to commutation of the Zeeman term with spin Hamiltonians \eqref{ham1}, \eqref{ham2} and to particular forms of spin representations \eqref{spinrep}, \eqref{hp}, and \eqref{drep}. Magnetic field reduces the gap in the spectrum of one of the branches, bringing the system to a quantum critical point. \section{Disorder modeling and technique} \label{method} We now turn to the systems considered above with finite concentration $c$ of defects. Here we discuss impurities which change only exchange coupling constants in corresponding Hamiltonians and which do not change the nature of the paramagnetic phase at $H=0$ (i.e., the ground state remains singlet). For instance, we do not consider defects below which weaken intradimer coupling constants so much that local magnetic moments arise on imperfect bonds. On the other hand, we do not assume below that deviation of the coupling constants on imperfect bonds from their values in pure systems is small. Two types of disorder can be distinguished: i) disorder in the intradimer exchange coupling constant $\cal J$ or in the value of the single-ion anisotropy $\cal D$ and ii) disorder in small exchange coupling constants $J_{ij}$ between spins from different dimers or spins on neighboring sites (in large-$\cal D$ systems). Hamiltonians of systems with defects are written as \begin{equation} \mathcal{H}=\mathcal{H}_2+V, \label{impHam} \end{equation} where $\mathcal{H}_2$ is given by Eqs.~\eqref{h2}, \eqref{h22}, \eqref{h23}, and \eqref{h24} and $V$ has the following form for disorder in $\cal J$ or $\cal D$ only \begin{equation} V = \sum_{\{n\}} u \mathbf{S}_{n,1} \cdot \mathbf{S}_{n,2}, \text{ or } V = \sum_{\{n\}} u \left(S^z_{n}\right)^2, \label{v} \end{equation} where the summation is taken over all imperfect bonds or sites and $u$ measures the deviation of $\cal J$ or $\cal D$ on imperfect bonds or sites from their values in pure systems. It is seen from Eqs.~\eqref{spinrep}, \eqref{hp} and \eqref{drep} that such a disorder effects only the chemical potential value on imperfect bonds or sites which is parametrized by the single parameter $u$. Thus, one obtains at $H<H_{c1}$ from Eqs.~\eqref{v} for one sort of particles ($\mathfrak a$-particles, for definiteness) \begin{equation} V = u \sum_{\{n\}}\mathfrak a^+_n \mathfrak a_n. \label{v1} \end{equation} \begin{figure} \noindent \hfil \includegraphics[scale=0.5]{SpinLadder1.eps} \includegraphics[scale=0.5]{AnisotropSys.eps} \hfil \caption{1D systems with imperfect bonds shown by dashed lines. (a) Spin-$\frac12$ ladder with dimers on rungs (shown in bold) with modified intradimer exchange constant $\cal J$ at rung 1 and modified values of exchange coupling constants between spins from dimer 1 and neighboring dimers 0 and 2. (b) Integer spin chain with modified value of the single-ion easy-plain anisotropy $\cal D$ at site 1 and modified value of exchange coupling constant at bonds 0-1 and 1-2.} \label{systems1D} \end{figure} Expressions for $V$ are cumbersome for disorder in $J_{ij}$ and we present them for 1D systems only which are shown in Fig.~\ref{systems1D}: \begin{equation} \begin{split} &V = u_1\sum_{\{n\}} \left(\mathbf{S}_{n,1} \cdot \mathbf{S}_{n+1,1}+ \mathbf{S}_{n,2} \cdot \mathbf{S}_{n+1,2} + \mathbf{S}_{n-1,1} \cdot \mathbf{S}_{n,1} + \mathbf{S}_{n-1,2} \cdot \mathbf{S}_{n,2}\right)\\ \text{ or } &V = u_1\sum_{\{n\}} \left(\mathbf{S}_{n} \cdot \mathbf{S}_{n+1}+ \mathbf{S}_{n-1} \cdot \mathbf{S}_{n} \right), \end{split} \end{equation} where the first equation is for the spin-$\frac12$ ladder (see Fig.~\ref{systems1D}(a)), the second one is for the integer spin chain (see Fig.~\ref{systems1D}(b)), $u_1$ measures the deviation of $J_{ij}$ on imperfect bonds from its value in pure systems. The part of the perturbation operators corresponding to $\mathfrak a$-particles has the form \begin{equation} V = \frac {a u_1}{2}\sum_{\{n\}} \left( \mathfrak a^+_n \mathfrak a_{n+1} + \mathfrak a^+_{n+1}\mathfrak a_n + \mathfrak a^+_{n-1} \mathfrak a_n + \mathfrak a^+_n \mathfrak a_{n-1} \right), \label{v2} \end{equation} where $a=1$ for spin ladder and $a=S(S+1)$ for integer-spin chains if $H=0$. We omit in Eq.~\eqref{v2} terms of the form $\mathfrak b^+_i \mathfrak a^+_j$, $\mathfrak b_i \mathfrak a_j$, $\mathfrak c^+_i \mathfrak c^+_j$, and $\mathfrak c_i \mathfrak c_j$, which arises at small $H$ and give corrections of the next order in $J_{ij}$. We start our discussion below with the disorder in $\cal J$ or $\cal D$ only. Then, we also add the disorder in $J_{ij}$ and discuss corresponding results for systems with two types of disorder. As it is usually done \cite{Vojta,Zhel13}, we assume that these two types of disorder are "coupled". For example, spins from an imperfect dimer are coupled to spins from neighboring dimers by imperfect bonds (see, e.g., Fig.~\ref{systems1D} for 1D systems). This assumption is quite natural because in real materials substitution of a non-magnetic atom usually changes all exchange coupling constants in its vicinity. \subsection{$T$-matrix approach} We use the conventional $T$-matrix approach (see, e.g., Refs.~\cite{Izyumov,white,wan,donia}) to find analytically corrections to quasiparticles spectra and density of states (DOS). Processes involving simultaneous scattering on more than a single impurity are omitted in this technique and all results are valid in the first order in $c$. One obtains for Green's functions of each mode in disordered systems \begin{equation} G(\mathbf{k},E)=\frac{1}{E-\varepsilon_{\mathbf{k}}-cT({\bf k},E)}, \label{green1} \end{equation} where $T({\bf k},E)$ is a quantity related to the $T$-matrix and $\varepsilon_{\mathbf{k}}$ is the pure system quasiparticle spectrum. Then, the translation invariance of systems is effectively restored in the first order in $c$ that allows using Green's functions of the form \eqref{green1} to analyze the spectra of these modes. \cite{Izyumov,white,wan} The quantity $T({\bf k},E)$ can be expressed via coordinate Green's functions of pure systems \begin{equation} G_{nm}(E)=\frac{1}{N} \sum_\mathbf{p} \frac{e^{i \mathbf{p}(\mathbf{R}_n-\mathbf{R}_m) }}{E-\varepsilon_{\mathbf{p}} - i0}, \label{Green1} \end{equation} where $N$ is the number of unit cells. For disorder in $\cal J$ or $\cal D$ only, $T({\bf k},E)$ does not depend on momentum having the form \begin{equation} T({\bf k},E)=\frac{u}{1- uG_{00}(E)}, \label{Wk1} \end{equation} where $u$ measures the deviation of $\cal J$ or $\cal D$ on imperfect bonds or sites from their values in pure systems (see Eq.~\eqref{v}). Spectrum of quasiparticles $E_\mathbf{k}$ and their damping $\gamma_\mathbf{k}$ are defined by poles of Green's function \eqref{green1} and they have the form in the first order in $c$ \begin{equation} E_\mathbf{k}=\varepsilon_{\mathbf{k}}+c \Re{(T({\bf k},E=\varepsilon_{\mathbf{k}}))}, \quad \gamma_\mathbf{k}=c \Im{(T({\bf k},E=\varepsilon_{\mathbf{k}}))}, \label{encor} \end{equation} where $\Re$ and $\Im$ denote real and imaginary parts, respectively. Eqs.~\eqref{encor} are written under assumption that the solution of the equation $E-\varepsilon_{\mathbf{k}}-cT({\bf k},E)=0$ at fixed $\bf k$ can be expanded as series in $c$ in which the first terms taken into account in Eqs.~\eqref{encor} are much larger than higher order terms. It can happen, however, that this is not the case for some $\bf k$. It would signify that diagrams of higher orders in $c$ have to be taken into account and Eqs.~\eqref{green1}--\eqref{encor} have to be reconsidered. As we obtain below, it is the situation that arises in considered systems for states in excitation band lying near its bottom and the top, where the following inequalities should hold for Eqs.~\eqref{encor} validity: \begin{eqnarray} \label{valid} &&|\varepsilon_{\mathbf{k}}-\Delta| \gg c|T({\bf k},\varepsilon_{\mathbf{k}})|,\\ \label{valid2} &&|\varepsilon_{\mathbf{k}}-\Delta-a|J_{\bf k_0}|| \gg c|T({\bf k},\varepsilon_{\mathbf{k}})|, \end{eqnarray} respectively. To analyze states near bands edges, we perform numerical calculations discussed below in details. It should be noted that invalidity of Eqs.~\eqref{green1}--\eqref{encor} and the necessity to go beyond the first order in $c$ is usually seen from an analysis similar to that just described. It happens sometimes that processes of multiple-defects scattering are important and their contributions (which are of higher orders in $c$) are much larger than the first order corrections. We demonstrate below that it is the situation which arises in 1D systems under discussion. Defects modify the system DOS \cite{Izyumov,white} $g(E)$. The general expression for $g(E)$ has the following form in the first order in $c$ in the case of disorder in $\cal J$ or $\cal D$ only: \begin{equation} g(E) = g_0(E) - c\frac{u^2 g_0(E) \Re (dG_{00}/dE)+u(1-u \Re (G_{00}(E)))dg_0/dE}{\left[ (1-u \Re (G_{00}(E)))^2 + (\pi u g_0(E))^2\right]}, \label{ds1} \end{equation} where $g_0(E)=\Im(G_{00}(E))/\pi$ is the pure system DOS and $G_{00}(E)$ is given by Eq.~\eqref{Green1} with $m=n=0$. It is seen from Eq.~\eqref{ds1} that the correction to DOS can have extrema when the following condition is satisfied: \begin{equation} 1-u \Re(G_{00}(E))=0. \label{ds2} \end{equation} It is well-known that in magnetically ordered phases solutions of equations similar to Eq.~\eqref{ds2} give positions of isolated levels (localized states) outside the excitation band (where $g_0(E)=0$) or virtual levels (resonances) inside the band. \cite{Izyumov,white} However, we find below that Eq.~\eqref{ds2} gives only positions of isolated impurity levels in the paramagnetic phases and all anomalies inside the band stem from derivatives in the numerator of the second term in Eq.~\eqref{ds1}. Imperfection in $J_{ij}$ can be taken into account in the same way although the corresponding analytical consideration is more technically involved. Some details on this point can be found in Appendix~\ref{append} devoted to 1D systems. Green's functions of propagating modes have the form \eqref{green1}, where $T({\bf k},E)$ does depend on momentum $\bf k$ and Green's functions \eqref{Green1} with $m\ne n$ also contribute to it. As a result expressions for $T({\bf k},E)$ and DOS are more cumbersome than Eqs.~\eqref{Wk1} and \eqref{ds1} and we do not present them here although the spectrum renormalization is given in the first order in $c$ by Eqs.~\eqref{encor} as before. \subsection{Numerical calculations} \label{numcalc} To confirm our analytical results and to reveal the nature of states near excitation bands edges, we perform numerical diagonalization of the one-particle sector of the bosonic Hamiltonians \eqref{impHam}, \eqref{v1} and \eqref{v2} for finite 1D and 2D systems with disorder. Thus, we find eigenvalues, eigenfunctions and DOS $\rho(\epsilon)$ of finite systems. Energy and damping of a propagating mode with momentum $\bf k$ are found using the Green's function definition \begin{equation} G({\bf k},t) = -i\langle vac |T\mathfrak a_{\bf k}(t)\mathfrak a^\dagger_{\bf k}(0) |vac\rangle = -i\langle {\bf k} | e^{-i{\cal H}t} |{\bf k}\rangle \theta(t) = -i\sum_\epsilon \left|\langle {\bf k} | \epsilon \rangle\right|^2e^{-i\epsilon t} \theta(t), \label{g} \end{equation} where $|vac\rangle$ is the ground state of the Hamiltonian $\cal H$ given by Eqs.~\eqref{impHam}, \eqref{v1} and \eqref{v2}, $|{\bf k}\rangle$ is the state with a particle having momentum $\bf k$ (plane wave), $\theta(t)$ is the Heaviside step function, and $|\epsilon \rangle$ is the eigenfunction of $\cal H$ corresponding to eigenvalue $\epsilon$. One can replace the summation on $\epsilon$ by integration inside the band and we have from Eq.~\eqref{g} \begin{equation} G({\bf k},\omega) = \int d\epsilon\frac{f({\bf k},\epsilon)}{\omega-\epsilon+i0} = \dashint d\epsilon\frac{f({\bf k},\epsilon)}{\omega-\epsilon}-i\pi f({\bf k},\omega), \label{g22} \end{equation} where $f({\bf k},\epsilon)=\rho(\epsilon)|\langle {\bf k} | \epsilon \rangle|^2$ can be found from the exact diagonalization results and $\dashint$ denotes the principal value of the integral. It follows from Eq.~\eqref{g22} that $f({\bf k},\omega)$ is related to the imaginary part of the Green's function $G({\bf k},\omega)$ which should have the Lorentzian shape for a well-defined propagating quasiparticle with momentum $\bf k$. Thus, one can obtain the energy and the damping of propagating excitations by fitting $f({\bf k},\omega)$ with the Lorentzian. To characterize quantitatively the spatial localization/delocalization of a state $\psi$ found by diagonalization, we calculate also the inverse participation ration (IPR): \begin{equation} \label{iprexpr} \mbox{IPR}(\psi) = \sum_n |\psi(n) |^4, \end{equation} where $n$ labels the lattice sites. IPR is of the order of the inverse number of sites occupied at state $\psi$. Then, IPR scales as $1/L^d$ for spatially extended states and it is equal to a constant for localized states, where $d$ is the system dimension. Exponential localization is characterized by ${\rm IPR}\propto 1/\xi^d$, where $\xi$ is of the order of localization length (see, e.g., Ref.~\cite{Vojta}). The number of sites in considered clusters vary from 400 to 15000. For each cluster, we perform an averaging over a large number of disorder realizations to find $f({\bf k},\epsilon)$, DOS and IPR. The number of disorder realizations vary from $10^5$ for the smallest clusters to 600 for the largest ones. We try both periodic and open boundary conditions which lead to the same results. Corrections to the quasiparticles energy and their damping are found by an extrapolation of numerical data for a number of finite size systems containing $L^d$ unit cells to thermodynamic limit using quadratic polynomials in $1/L$. In particular, Fig.~\ref{1dspec} presented in the next section for 1D systems is build using 250 momentum values. Extrapolations in planes \ref{1dspec}(a) and \ref{1dspec}(b) are carried out using $L=1024$, 2048, 4096, and 8192. $f({\bf k},\epsilon)$ shown in insets are calculated for $L=6144$. In planes \ref{1dspec}(c) and \ref{1dspec}(d), clusters with $L=500$, 1000, 2000, and 3000 are used for extrapolations and insets show results for $L=3000$. \section{Disordered systems} \label{disordersec} \subsection{1D systems} \subsubsection{$T$-matrix approach} Calculations are particularly simple in 1D systems with defects which are depicted in Fig.~\ref{systems1D}. Taking into account the exchange coupling between nearest neighbors only, one has for the bare spectrum (cf. Eq.~\eqref{spec}) \begin{equation} \varepsilon_{\bf k} = \Delta+a|J|+aJ\cos k, \label{1dspec2} \end{equation} where $\Delta = 1-a|J|$ at $H=0$ and the spectrum minimum is located at $k=k_0$, where $k_0=\pi$ and 0 for $J>0$ and $J<0$, respectively. We obtain after simple integration in Eq.~\eqref{Green1} \begin{equation} G_{00}(E)= \begin{cases} \frac{1}{\sqrt{(E-\Delta-a|J|)^2-a^2J^2}}, & E>\Delta+2a|J|, \\ \frac{i}{\sqrt{a^2J^2-(E-\Delta-a|J|)^2}}, & \Delta< E\le \Delta+2a|J|, \\ -\frac{1}{\sqrt{(E-\Delta-a|J|)^2-a^2J^2}}, & E<\Delta. \label{g001d} \end{cases} \end{equation} Using the second line in Eq.~\eqref{g001d}, one has $ G_{00}(E=\varepsilon_{k})=i/|aJ \sin{k}| $ and we obtain for the spectrum and the damping from Eqs.~\eqref{Wk1} and \eqref{encor} in the case of disorder in $\cal J$ or $D$ only \begin{equation} E_k = \Delta+a|J|+aJ\cos k + c \frac{ua^2J^2 \sin^2 k}{a^2J^2 \sin^2 k+u^2}, \qquad \gamma_k=c \frac{u^2 a|J \sin{k}|}{a^2J^2 \sin^2 k+ u^2}. \label{spec1d} \end{equation} Let us discuss the neighborhood of the spectrum minimum, where it has the form \eqref{spec0}. It is seen from Eqs.~\eqref{spec1d} that there are two regimes at $\kappa=|k-k_0|\ll1$: \begin{equation} \begin{array}{llll} E_k &= \Delta + \left(\frac{a|J|}{2} + c\frac{a^2J^2}{u} \right)\kappa^2, \qquad &\gamma_k = ca|J|\kappa, \quad &\mbox{if}\quad \kappa\ll \min\{1,|u/aJ|\},\\ E_k &= \Delta + cu + \frac{a|J|}{2} \kappa^2, \qquad &\gamma_k = c\frac{u^2}{a|J|\kappa}, \quad &\mbox{if}\quad 1\gg\kappa\gg |u/aJ|. \label{E1d} \end{array} \end{equation} The range of Eqs.~\eqref{E1d} validity given by Eq.~\eqref{valid} reads \begin{equation} \begin{array}{ll} \kappa\gg c,\qquad &\text{if}\quad |u|\gg ca|J|,\\ \kappa\gg \sqrt{c\left|\frac{u}{aJ}\right|}, \qquad &\text{if}\quad |u|\ll ca|J|. \end{array} \label{val1d} \end{equation} One is lead from Eq.~\eqref{valid2} to the same range of the analytical approach validity near the spectrum maximum (in this case $\kappa$ measures a deviation of momentum from the value at which the spectrum has the maximum). Correction to the quasiparticle energy and its damping given by Eqs.~\eqref{spec1d} are plotted in Fig.~\ref{1dspec} for particular parameters values. The corresponding numerical results are also shown in Fig.~\ref{1dspec} which are discussed below. \begin{figure} \noindent \includegraphics[scale=0.4]{1d-energy2.eps} \includegraphics[scale=0.4]{1d-damping2.eps} \includegraphics[scale=0.4]{1d-energy.eps} \includegraphics[scale=0.4]{1d-damping.eps} \hfil \caption{ (Color online.) Correction to quasiparticles energy and their damping in 1D systems given by Eqs.~\eqref{spec1d} and found numerically for disorder in $\cal J$ or $\cal D$ only (the extrapolation is carried out of numerical data for finite systems containing $L$ unit cells to thermodynamical limit as it is explained in Sec.~\ref{numcalc}). Shaded regions mark areas in which the imaginary parts of one-particle Green's functions $\chi''(k,\omega)$ found numerically using Eq.~\eqref{g22} do not have a Lorentzian shape and in which our analytical results are invalid (i.e., inequalities \eqref{val1d} do not hold). Insets in planes (b) and (d) show $\chi''(k,\omega)$ for some fixed momenta, where solid lines represent results of data fitting by Lorentzians. Most pronounced anomalies in numerical data for ``stronger'' impurities (planes (c) and (d)) are interpreted as a result of coherent scattering by defects of quasiparticles with momenta denoted by vertical lines (see the text). } \label{1dspec} \end{figure} DOS of pure systems $g_0(E)$ is equal to $G_{00}(E)/i\pi=\frac1\pi(a^2J^2-(E-\Delta-a|J|)^2)^{-1/2}$ inside the band. One concludes from Eq.~\eqref{ds1} that defects do not lead to noticeable corrections to DOS in the range of the analytical approach validity determined by Eqs.~\eqref{val1d}. Outside the band, where $G_{00}(E)$ is real, an isolated impurity level appears above or below the band depending on the sign of $u$. One has from Eqs.~\eqref{ds1}, \eqref{ds2}, and \eqref{g001d} for $E$ lying outside the band (see Fig.~\ref{dos1}(a)) \begin{equation} g(E) = c \delta(E-E_d), \qquad E_d = \Delta+a|J|+{\rm sign}(u) \sqrt{a^2J^2+u^2}. \label{imp1d} \end{equation} Multiple-impurities scattering processes, which are not taken into account in the first order in $c$, turn this isolated level into a narrow impurity band. Eqs.~\eqref{imp1d} are in accordance with the corresponding result of Ref.~\cite{Arlego} devoted mainly to DOS in disordered spin-$\frac12$ ladders. \begin{figure} \noindent \includegraphics[scale=0.4]{DOS1D.eps} \includegraphics[scale=0.4]{DOS1D2.eps} \hfil \caption{(Color online.) DOS of 1D systems with disorder in $\cal J$ or $\cal D$ only, where $\delta E=E-\Delta-a|J|$. DOS in the $T$-matrix approach is given by Eq.~\eqref{ds1} (it is almost indistinguishable on the plots from the pure system DOS). Most pronounced anomalies in numerical data found for $L=3000$ are interpreted as a result of coherent scattering by defects of quasiparticles with energies denoted by vertical lines (see the text).} \label{dos1} \end{figure} Imperfection in the small exchange coupling $J_{ij}$ is considered in detail in Appendix~\ref{append}. Eq.~\eqref{w1d2} is derived there for $T(k,E)$ that gives for the spectrum using Eqs.~\eqref{encor} (cf.\ Eqs.~\eqref{spec1d}) \begin{eqnarray} E_k &=& \Delta+a|J|+aJ\cos k + c\frac{\left( 1+\frac{u_1}{J}\right)^2\left(u+2 a u_1\cos k + a u^2_1\cos k/J\right)a^2 J^2 \sin^2 k}{\left( 1+ \frac{u_1}{J}\right)^4a^2 J^2 \sin^2 k + \left(u+2 a u_1\cos k +a u^2_1\cos k/J\right)^2},\nonumber\\ \gamma_k &=& ca|J \sin k| \frac{\left(u+2 a u_1\cos k + a u^2_1\cos k/J\right)^2}{\left( 1+ \frac{u_1}{J}\right)^4 a^2 J^2 \sin^2 k + \left(u+2 a u_1\cos k +a u^2_1\cos k/J\right)^2}. \label{spec1d22} \end{eqnarray} If $|u|J|-2 a u_1 J- au^2_1|\gg a(J+u_1)^2 |\sin k|$, the spectrum has the form near its minimum (cf.\ Eqs.~\eqref{E1d}) \begin{equation} E_k=\Delta+\left(\frac{a|J|}{2} + c\frac{a^2J^2(1+u_1/J)^2}{u-2 a u_1 J/|J|- au^2_1/|J|} \right)\kappa^2, \qquad \gamma_k=ca|J|\kappa. \label{spec1d2} \end{equation} We point out also the reduction of the spectrum renormalization by two sorts of disorder when $u|J|-2 a u_1 J- au^2_1\approx0$ and $|1+u_1/J|\sim1$. One obtains in this case from Eqs.~\eqref{spec1d22} $|E_k-\varepsilon_k|\sim c|J|\kappa^2\ll\varepsilon_k-\Delta$ and $\gamma_k\sim c|J|\kappa^3\ll\varepsilon_k-\Delta$. DOS in 1D systems with two sorts of disorder is also considered in Appendix~\ref{append}. It is shown there, in particular, that there are no isolated impurity levels at $a|u_1|\gg|u|$ if $-2<u_1/J<0$, whereas one level above and one level below the band arise if $u_1/J$ lies outside this interval (in accordance with Ref.~\cite{Arlego}). It should be noted that Eqs.~\eqref{g001d}, \eqref{spec1d}, and \eqref{spec1d22} are derived using the particular form of the spectrum \eqref{1dspec2}. One would lead to different results for gapped phases with another spectrum. In contrast, Eqs.~\eqref{E1d}--\eqref{val1d} and \eqref{spec1d2} are more universal because they can be obtained using the general form of the spectrum \eqref{spec0} near its minimum (or maximum) and the form of the impurity interaction \eqref{v1} and \eqref{v2} (the combination $a|J|$ in these expressions stems from the factor in the expression \eqref{spec0} for the spectrum and $au_1$ originates from $V$ given by Eq.~\eqref{v2}). This is due to the fact that small $\kap_p$ give the main contribution to $G_{nm}$ in Eq.~\eqref{Green1}, where $\kap_p$ is the deviation of $\bf p$ from the momentum at which the spectrum has minimum (or maximum). \subsubsection{Numerical results} \label{num1d} Our numerical results for the quasiparticle energy, damping and DOS are also presented in Figs.~\ref{1dspec} and \ref{dos1} (for the disorder in $\cal J$ or $\cal D$ only). As it is seen, they are in good agreement with analytical findings in the range of the analytical approach validity \eqref{val1d} except for some points near which upward and downward spikes appear. Amplitudes of these spikes rise as $|u|$ and/or $c$ increase. As the $T$-matrix approach does not show such anomalies, we attribute them to resonances in multiple scattering on defects which are not taken into account in our analytical consideration and which are effects of higher order in $c$. The origin of these resonances can be understood qualitatively by noting that elementary excitations of a pure chain with momenta $k=m\pi/n$ and $k=\pi-m\pi/n$, where $m<n$ are integers, scatter coherently by defects which are $rn$ sites apart, where $r$ is integer. If the renormalized spectrum $E_k$ differs noticeably from the bare spectrum $\varepsilon_k$ given by Eq.~\eqref{1dspec2}, positions of anomalies shift a little due to the fact that an excitation with energy $E_k$ produces excitations with the same energy $\varepsilon_p=E_k$ as a result of scattering on defects which interfere coherently if $p=m\pi/n$. Positions of resonances found in this way are denoted in Figs.~\ref{1dspec}(c), \ref{1dspec}(d) and \ref{dos1} by vertical lines which mark accurately location of anomalies in numerical data (momenta $p$ are also depicted in Fig.~\ref{dos1}(b) near corresponding vertical lines). The imaginary part of the one-particle Green's function $\chi''(k,\omega)$ is shown for some momenta in insets of Figs.~\ref{1dspec}(b) and \ref{1dspec}(d) which have been found numerically as it is described above. These insets illustrate our finding that $\chi''(k,\omega)$ has the Lorentzian shape for not too strong impurities in the range of the analytical approach validity. Upon $u$ and/or $c$ increasing, amplitudes of anomalies rise and the form of peaks in $\chi''(k,\omega)$ in the vicinity of corresponding $k$ bears little resemblance to a Lorentzian for large enough $u$. The resonant scattering becomes strong enough and our analytical results are completely invalid when $c|u/aJ|\agt1$ (see Fig.~\ref{dos1}(b)). Peaks in $\chi''(k,\omega)$ have non-Lorentzian shapes near the band edges for all $u$ and $c\ll1$ (see Fig.~\ref{1dspec}(b) for illustration). Areas in $k$-space with non-Lorentzian peaks are shaded in Fig.~\ref{1dspec} that illustrates our results for $|u/aJ|\sim1$ when anomalies inside the band are not too large. These areas widths are in accordance with our estimations \eqref{val1d} of regions sizes in which analytical results are valid. Our analysis of IPR defined by Eq.~\eqref{iprexpr} demonstrates that all the states in the band are localized for any $u$ and $c$. This is illustrated by Fig.~\ref{iprfig}(a) drawn for one state inside the band far from its edges. Data are averaged over disorder realizations and the mean square deviation of IPR from its mean value is shown in Fig.~\ref{iprfig}(b). Interestingly, states inside the band far from its edges can combine the localization and properties of short-wavelength wavepacket (if the resonant scattering is not too strong). This situation holds even in the limit $u\to+\infty$ and $u_1\to-J$, when defects break a system to pieces of mean length $1/c$. As is seen from Eqs.~\eqref{spec1d}--\eqref{val1d} and \eqref{spec1d22}, $c$-corrections vanish in expressions for short-wavelength quasiparticles energies but a finite damping remains that reflects a finite lifetime of wavepackets excited in such a broken system. \begin{figure} \noindent \includegraphics[scale=0.4]{ipr.eps} \hfil \caption{(Color online.) The inverse participation ration (IPR) given by Eq.~\eqref{iprexpr} averaged over disorder realizations as it is described in Sec.~\ref{numcalc}. $\sigma(\rm IPR)$ is the mean square deviation of IPR from its mean value. (a), (b) and (c), (d) slides are for particular states inside the band far from its edges in 1D and 2D systems, respectively. The states energies $\delta E$ are measured from the band center (see Figs.~\ref{dos1}(a) and \ref{dos2}, correspondingly). Insets show histograms of IPR distributions in disorder realizations. } \label{iprfig} \end{figure} \subsection{2D systems} \begin{figure} \noindent \includegraphics[scale=0.7]{Bilayer2.eps} \hfil \caption{Spin-1/2 dimerized bilayer with imperfect bonds. Notations are the same as in Fig.~\ref{systems1D}(a).} \label{SB1} \end{figure} We turn to 2D systems with the exchange coupling between nearest neighbors (see Fig.~\ref{SB1} for 2D dimer system) which spectrum has the form \begin{equation} \varepsilon_{\bf k} = \Delta + 2a|J| + aJ(\cos k_x + \cos k_y), \end{equation} where $\Delta = 1-2a|J|$ at $H=0$. One obtains taking integral in Eq.~\eqref{Green1} for energies $E>\Delta+4a|J|$ lying outside the band \begin{equation} G_{00}(E)=\frac{2}{\pi(E-\Delta-2a|J|)} K\left(\frac{4a^2J^2}{(E-\Delta-2a|J|)^2} \right), \label{Green2D} \end{equation} where $K(k)=\int^{\pi/2}_0 \frac{d \theta}{\sqrt{1-k^2 \sin^2 \theta}}$ is the complete elliptic integral of the first kind. For energies inside the band, $E>\Delta+2a|J|$, the result can be represented in the form \begin{equation} G_{00}(E)=\frac{1}{\pi a|J|} \left[ \frac{1}{\cos \psi} F\left(\frac{\pi}{2}-\psi, \frac{1}{\cos \psi} \right) + \frac{i}{\sin \psi} F\left(\psi, \frac{1}{\sin \psi} \right) \right], \label{G02D} \end{equation} where $\psi=\arccos \left( \frac{E-\Delta-2a|J|}{2a|J|}\right)$, $F(\phi, k)=\int^{\phi}_0 \frac{d \theta}{\sqrt{1-k^2 \sin^2 \theta}}$ is incomplete elliptic integral of the first kind, and both of the elliptic functions are real. For other $E$ values, $G_{00}(E)$ can be easily found from Eqs.~\eqref{Green2D} and \eqref{G02D} by using the fact that its real and imaginary parts are antisymmetric and symmetric functions with respect to the point $E=\Delta+2a|J|$, correspondingly. Eq.~\eqref{G02D} can be simplified considerably at $E=\varepsilon_{\mathbf{k}}$ near the spectrum minimum ($\kappa=|k-k_0|\ll1$): \begin{equation} G_{00}(\varepsilon_{\mathbf{k}})\approx \frac{1}{\pi a|J|} \ln\frac{\kappa}{b_{21}} + \frac{i}{2a|J|}, \label{g002} \end{equation} where $b_{21}=\exp{(C_{20})}=2^{5/2}$, $C_{20}$ is a model dependent coefficient, \begin{equation} C_{20}=\pi a |J| \lim_{{\bf k}_1\rightarrow {\bf k}_0} \left( \frac{1}{(2\pi)^2} \int_{\Omega} \frac{d^2k}{\varepsilon_{\bf k} - \varepsilon_{{\bf k}_1}} + \ln k_1\right). \label{C_20} \end{equation} Using Eqs.~\eqref{Wk1}, \eqref{encor} and \eqref{g002} we have in the vicinity of the spectrum minimum for the disorder in $\cal J$ or $\cal D$ only \begin{equation} E_{\bf k} = \Delta+\frac{a|J|}{2}\kappa^2+c \frac{\pi a|J|(\pi a|J|-u\ln(\kappa/b_{21}))u}{(\pi a|J|-u\ln(\kappa/b_{21}))^2+(\pi u/2)^2}, \quad \gamma_{\bf k} = c \frac{\pi^2 }{2}\frac{a|J|u^2}{(\pi a|J|-u\ln(\kappa/b_{21}))^2+(\pi u/2)^2}. \label{2dcorbot} \end{equation} One obtains in the same way for the spectrum near the top of the band (i.e., near the spectrum maximum) \begin{equation} E_{\bf k} = \Delta+4a|J|-\frac{a|J|}{2}\kappa^2+c \frac{u\pi a|J|(\pi a|J|+u\ln(\kappa/b_{21}))}{(\pi a|J|+u\ln(\kappa/b_{21}))^2+(\pi u/2)^2}, \quad \gamma_{\bf k} = c \frac{\pi^2 }{2}\frac{u^2a|J|}{(\pi a|J|+u\ln(\kappa/b_{21}))^2+(\pi u/2)^2}. \label{2dcortop} \end{equation} The range of Eqs.~\eqref{2dcorbot} and \eqref{2dcortop} validity is written as \begin{equation} \begin{array}{cc} \kappa\gg\sqrt c, &\mbox{if } |u|\gg a|J|,\\ \kappa\gg\sqrt{c\left|\frac{u}{aJ}\right|}, &\mbox{if } |u|\ll a|J|, \end{array} \label{val2d} \end{equation} where $\kappa$ measures a deviation of momentum from the values at which the spectrum has maximum or minimum. Notice that all corrections to the spectrum depend weakly on momenta in the range of the results validity: $|E_{\bf k}-\varepsilon_{\bf k}|\sim c$ and $\gamma_{\bf k}\sim c$. The analytical approach is not valid for states near the top and the bottom of the band due to localization of excitations that is illustrated by Fig.~\ref{loc2d} found numerically for a single disorder realization. We have also observed that IPR$\propto1/L^{\alpha d}$ for states inside the band far from its edges, where $\alpha<1$ (see Fig.~\ref{iprfig}(c) and \ref{iprfig}(d)). Although this behavior differs from that of ordinary propagating excitations ($1/L^d$), the localization length $\xi\propto L^\alpha$ is infinite in the thermodynamic limit in the considered 2D systems. \begin{figure} \noindent \hfil \includegraphics[scale=0.3]{2d-disorder.eps} \includegraphics[scale=0.3]{s0.eps} \includegraphics[scale=0.3]{s10.eps} \includegraphics[scale=0.3]{s100.eps} \includegraphics[scale=0.3]{s1000.eps} \includegraphics[scale=0.3]{s10001.eps} \includegraphics[scale=0.3]{s10002.eps} \includegraphics[scale=0.3]{s102.eps} \includegraphics[scale=0.3]{simp.eps} \includegraphics[scale=0.3]{simp2.eps} \includegraphics[scale=0.27]{color_scale.eps} \hfil \caption{(Color online.) (a) Spatial distribution of defects with $c=0.1$, $u=3a|J|$, and $u_1=0$ in 2D system with the size $120\times120$ unit cells. (b)--(j) Numerically found color plots of wave functions amplitudes for the Hamiltonian of this disordered system which correspond to indicated eigenvalues $E$. Panels (b)--(e) give a picture of energy levels near the band bottom, panels (f)--(h) illustrate the band top, and panels (i) and (j) describe the impurity band corresponding to the localized level in the first order in $c$ (see also Fig.~\ref{dos2} for DOS found for the same parameters). All states in the impurity band are localized. $\Delta$ is the gap value for the particular disorder realization.} \label{loc2d} \end{figure} Defects impact on DOS is described by Eqs.~\eqref{ds1} and \eqref{ds2} which are difficult to treat analytically in 2D systems. As in 1D systems, Eq.~\eqref{ds2} has a solution at any finite $u$ outside the band, so that an isolated impurity level arises above and below the band for positive and negative $u$, respectively. The largest corrections to DOS inside the band appear near the bottom, the center and the top of the band which stem from singular derivatives in the numerator of the second term in Eq.~\eqref{ds1}. Due to these large corrections, the $T$-matrix approach does not work in these regions. These results are illustrated by Fig.~\ref{dos2} which demonstrates, in particular, our finding that in contrast to 1D systems there are no anomalies in spectrum corrections and DOS related to multiple-defects scattering processes (cf.\ Fig.~\ref{dos1}(b)). The numerical analysis of wave-functions shows that states around the anomaly at the band center remain propagating. \begin{figure} \noindent \hfil \includegraphics[scale=0.45]{DOS2D.eps} \hfil \caption{(Color online.) DOS of 2D systems with disorder in $\cal J$ or $\cal D$ only, where $\delta E=E-\Delta-2a|J|$, $c=0.1$, $u=3a|J|$, and $u_1=0$ (cf.\ Fig.~\ref{dos1}). Numerical results are obtained for the system size $100\times100$ unit cells. } \label{dos2} \end{figure} Taking into account disorder in $J_{ij}$ and performing calculations similar to those presented in Appendix~\ref{append} for 1D systems, we obtain corrections to quasiparticles energy and their damping which are cumbersome for arbitrary $\bf k$. However, these results turn out to be a simple modifications of Eqs.~\eqref{2dcorbot} in the vicinity of the spectrum minimum ($\kappa=|k-k_0|\ll1$): \begin{eqnarray} E_{\bf k} &=& \Delta+\frac{a|J|}{2}\kappa^2+c \frac{\pi a |J|u'\left(\pi a|J|-u'\ln(\kappa/b_{21})+\left(b_{23} a\frac{J}{|J|}u_1+b_{24} u''\right)\right)}{\left(\pi a|J|-u'\ln(\kappa/b_{21})+\left(b_{23} a\frac{J}{|J|}u_1+b_{24} u''\right)\right)^2+\frac{\pi^2}{4}u^{\prime2}},\nonumber\\ \gamma_{\bf k} &=& c \frac{\pi^2}{2} \frac{a |J| u^{\prime2}}{\left(\pi a|J|-u'\ln(\kappa/b_{21})+\left(b_{23} a\frac{J}{|J|}u_1+b_{24} u''\right)\right)^2+\frac{\pi^2}{4}u^{\prime2}}, \label{spec2dbot} \end{eqnarray} where $u'=u-4au_1J/|J|-b_{22} au^2_1/|J|$, $u''=au^2_1/|J|$, \begin{eqnarray} b_{22}&=&(5C_{20}-2C_{211}-C_{22}-8C_{21})/\pi=1.44, \nonumber \\ b_{23}&=&C_{20}-C_{21}=1.57, \\ b_{24}&=&C_{20}b_{22}-(C^2_{20}-2C_{211}C_{20}-C_{22}C_{20}-4C^2_{21})/\pi=4.79, \nonumber \end{eqnarray} where $C_{21}, C_{22}, C_{211}$ are model dependent coefficients as \eqref{C_20}, \begin{eqnarray} C_{21}&=&\pi a |J| \lim_{{\bf k}_1\rightarrow {\bf k}_0} \left(- \frac{1}{(2\pi)^2} \int_{\Omega} \frac{d^2k \cos k_x}{\varepsilon_{\bf k} - \varepsilon_{{\bf k}_1}} + \ln k_1\right), \nonumber\\ C_{22}&=&-\pi a |J| \lim_{{\bf k}_1\rightarrow {\bf k}_0} \left( \frac{1}{(2\pi)^2} \int_{\Omega} \frac{d^2k \cos 2k_x}{\varepsilon_{\bf k} - \varepsilon_{{\bf k}_1}} + \ln k_1\right), \\ C_{211}&=&-\pi a |J| \lim_{{\bf k}_1\rightarrow {\bf k}_0} \left( \frac{1}{(2\pi)^2} \int_{\Omega} \frac{d^2k \cos k_x \cos k_y}{\varepsilon_{\bf k} - \varepsilon_{{\bf k}_1}} + \ln k_1\right). \nonumber \end{eqnarray} We lead to the following expressions near the top of the band which resemble Eqs.~\eqref{2dcortop}: \begin{eqnarray} E_{\bf k} &=& \Delta+4a|J|-\frac{a|J|}{2}\kappa^2 + c \frac{\pi a |J|u'\left(\pi a|J|+u'\ln(\kappa/b_{21})+\left(b_{23} a\frac{J}{|J|}u_1+b_{24} u''\right)\right)}{\left(\pi a|J|+u'\ln(\kappa/b_{21})+\left(b_{23} a\frac{J}{|J|}u_1+b_{24} u''\right)\right)^2+\frac{\pi^2}{4}u^{\prime2}},\nonumber\\ \gamma_{\bf k} &=& c \frac{\pi^2}{2} \frac{a |J| u^{\prime2}}{\left(\pi a|J|+u'\ln(\kappa/b_{21})+\left(b_{23} a\frac{J}{|J|}u_1+b_{24} u''\right)\right)^2+\frac{\pi^2}{4}u^{\prime2}}, \label{spec2dtop} \end{eqnarray} where now $u'=u+4au_1J/|J|+b_{22} au^2_1/|J|$ and $u''=au^2_1/|J|$. The weak dependence of corrections to the spectrum on momentum remains in the case of two types of disorder. It is seen from Eqs.~\eqref{spec2dbot} and \eqref{spec2dtop} that similar to 1D systems a mutual reduction of contributions from two sorts of disorder arises at $u|J| \approx 4au_1J+b_{22}au^2_1$ near the band bottom and at $u|J| \approx -4au_1J-b_{22}au^2_1$ near its top. Analysis of DOS shows that similar to 1D systems the disorder in $J_{ij}$ only leads to one impurity level above the band and one impurity level below it if $u_1$ lies outside the interval $-2<u_1/J<0$ and there are no isolated impurity levels for $u_1$ lying inside this interval. Similar to 1D systems, one leads to the same results \eqref{2dcorbot}--\eqref{val2d} and \eqref{spec2dbot}, \eqref{spec2dtop} using the general form of the spectrum \eqref{spec0} near its minimum (or maximum) because mainly small $\kap_p$ contribute to Green's functions $G_{mn}$ at small $\kap$. Model-dependent quantities in these expressions which depend on the form of the spectrum at $\kap_p\sim1$ are constants $b$. They are of the order of unity. \subsection{3D systems} \begin{figure} \noindent \includegraphics[scale=0.4]{3D.eps} \hfil \caption{3D spin-$\frac12$ dimer system with imperfect bonds. Notations are the same as in Figs.~\ref{systems1D}(a) and \ref{SB1}.} \label{3D} \end{figure} 3D spin-$\frac12$ dimer system under discussion is shown in Fig.~\ref{3D}. For the cubic lattice with interaction between nearest spins, the spectrum has the form \begin{equation} \varepsilon_{\mathbf{k}} = \Delta + 3a|J| + aJ(\cos k_x + \cos k_y + \cos k_z), \label{spec03d} \end{equation} where $\Delta = 1-3a|J|$ at $H=0$. Green's function \eqref{Green1} can be represented as follows: \begin{equation} G_{00}(E)=\frac{1}{\pi} \int^\pi_0 dz G^{(2D)}_{00}(E-a|J|-aJ \cos z), \label{Gr3D} \end{equation} where $G^{(2D)}_{00}$ is the Green's function \eqref{Green2D} for 2D systems. Eq.~\eqref{Gr3D} has the following form at $E=\varepsilon_{\mathbf{k}}$ near the spectrum minimum ($\kappa=|k-k_0|\ll1$): \begin{equation} G_{00}(\varepsilon_{\mathbf{k}})\approx -\frac{1}{b_{31}a|J|} + i \frac{\kappa}{2 \pi a|J|}, \label{g003d} \end{equation} where $b_{31}=1/C_{30}=2$, $C_{30}$ is a model dependent coefficient, \begin{equation} C_{30}=\frac{a|J|}{(2\pi)^3} \int \frac{d^3k}{\varepsilon_{\bf k} - \varepsilon_{{\bf k}_0}}. \label{C_30} \end{equation} Using Eqs.~\eqref{Wk1}, \eqref{encor} and \eqref{g003d}, we obtain for the spectrum near its minimum in the case of disorder in $\cal J$ or $\cal D$ only \begin{equation} E_{\bf k} = \Delta + \frac{a|J|}{2}\kappa^2 + c \frac{b_{31}a|J|u}{u + b_{31}a |J|}, \qquad \gamma_{\bf k} = c \kappa \frac 1\pi \frac{b_{31}^2a|J|u^2}{\left( u + b_{31}a |J|\right)^2}. \label{spec3dbot} \end{equation} It is seen from Eqs.~\eqref{spec3dbot} that the quasiparticle energy acquires a small correction and $\gamma_{\bf k}\sim c\kappa$ if $|u+b_{31}a |J||\gg |u|\kappa$. However, the damping enhances greatly, $\gamma_{\bf k}\sim c/\kappa$, if $|u+b_{31}a |J||\ll |u|$ that signifies an appearance of a resonant scattering by defects in the first order in $c$. In the vicinity of the spectrum maximum, one obtains the following results (cf.\ Eqs.~\eqref{spec3dbot}): \begin{equation} E_{\bf k} = \Delta + 6 a |J| - a\frac{|J|}{2}\kappa^2 - c \frac{b_{31}a|J|u}{u-b_{31}a |J|}, \qquad \gamma_{\bf k} = c \kappa \frac 1\pi \frac{b_{31}^2a|J|u^2}{\left( u-b_{31}a |J|\right)^2}. \label{spec3dtop} \end{equation} The resonant scattering takes place in this case if $|u-b_{31}a |J||\ll |u|$. If conditions $|u \pm b_{31}a |J||\ll |u|$ are not satisfied, the range of Eqs.~\eqref{spec3dbot} and \eqref{spec3dtop} validity is given by inequality $\kappa \gg c$. \begin{figure} \noindent \hfil \includegraphics[scale=0.4]{DOS3D.eps} \hfil \caption{DOS of 3D systems, where $\delta E=E-\Delta-3a|J|$, $c=0.1$, $u=3a|J|$, and $u_1=0$. Solid and dashed lines are for pure and discorded systems, respectively.} \label{dos3} \end{figure} Similar to 2D systems, Eqs.~\eqref{g003d}--\eqref{spec3dtop} are valid in other gapped models in which the spectrum differs from \eqref{spec03d} but depends quadratically on the momentum near its minimum and maximum. The model dependent constant $b_{31}$ is of the order of unity in this case. Effect of defects on DOS is illustrated by Fig.~\ref{dos3}. At $|u|<2a|J|$, there are no solutions of Eq.~\eqref{ds2} and there are no isolated impurity levels outside the band. If $|u|$ is large enough, $|u|>2a|J|$, the system has a localized level above or below the band for $u>0$ and $u<0$, respectively. Large corrections to DOS inside the band appear near its top and the bottom as well as at $E=\Delta+3a|J|\pm a|J|$ (see Fig.~\ref{dos3}) which stem from derivatives in the numerator of the second term in Eq.~\eqref{ds1}. The results obtained in the first order in $c$ are not valid near these anomalies. Taking into account the disorder in $J_{ij}$, one obtains for the spectrum near the band bottom (cf.\ Eqs.~\eqref{spec3dbot}) \begin{eqnarray} \label{spec3dbot2} E_{\bf k} &=&\Delta+\frac{a|J|}{2}\kappa^2+cu' a|J|, \quad \gamma_{\bf k}=c\kappa\frac{u^{\prime2}}{2\pi}a|J|,\\ \mbox{where } u' &=& b_{31}\frac{u-6au_1J/|J|-b_{32} au^2_1/|J|}{u+b_{31}a|J|-b_{33}au_1J/|J|-b_{34}au^2_1/|J|},\nonumber \end{eqnarray} \begin{eqnarray} b_{32}&=&(21 C_{30}+3 C_{32}+12 C_{311}-36C_{31}) /2=3, \nonumber \\ b_{33}&=&6 C_{31}/C_{30}=2, \\ b_{34}&=&3(C^2_{30}+4C_{311}C_{30}+C_{30}C_{32}-6 C^2_{31})=1, \nonumber \end{eqnarray} where $C_{31},C_{32},C_{311}$ are model dependent constants as \eqref{C_30}, \begin{eqnarray} C_{31}&=&\frac{a|J|}{(2\pi)^3} \int \frac{d^3k \cos k_x}{\varepsilon_{{\bf k}_0} - \varepsilon_{\bf k}}, \nonumber\\ C_{32}&=&\frac{a|J|}{(2\pi)^3} \int \frac{d^3k \cos 2k_x}{\varepsilon_{\bf k} - \varepsilon_{{\bf k}_0}}, \\ C_{311}&=&\frac{a|J|}{(2\pi)^3} \int \frac{d^3k \cos k_x \cos k_y}{\varepsilon_{\bf k} - \varepsilon_{{\bf k}_0}}. \nonumber \end{eqnarray} We have near the spectrum maximum (cf.\ Eqs.~\eqref{spec3dtop}) \begin{eqnarray} \label{spec3dtop2} E_{\bf k} &=& \Delta+6a|J|-\frac{a|J|}{2}\kappa^2 - cu'a|J|, \quad \gamma_{\bf k}=c\kappa\frac{u^{\prime2}}{2\pi}a|J|,\\ \mbox{where } u' &=& b_{31}\frac{u+6au_1J/|J|+b_{32} au^2_1/|J|}{u-b_{31}a|J|+b_{33}au_1J/|J|+b_{34}au^2_1/|J|}.\nonumber \end{eqnarray} Similar to lower dimensions considered above, the phenomenon of corrections compensation from two types of disorder arises in 3D systems as well: all corrections vanish at $u|J|=6au_1J+3au^2_1$ and $u|J|=-6au_1J-3au^2_1$ near the spectrum minimum and maximum, respectively. For disorder in $J_{ij}$ only, analysis of DOS shows that similar to 1D and 2D systems one impurity level above the band and one impurity level below it appear if $u_1$ lies outside the interval $-2.75<u_1/J<0.75$ and there are no isolated impurity levels for $u_1$ lying inside this interval that is wider in 3D systems compared to 1D and 2D ones. \section{Summary and conclusion} \label{sum} To summarize, we develop a theory based on the $T$-matrix approach which describes gapped phases in 1D, 2D, and 3D spin systems with bond disorder and with weakly interacting bosonic elementary excitations. Low-field paramagnetic and high-field fully saturated phases in dimerized spin-$\frac12$ magnets and integer-spin systems with large single-ion easy-plane anisotropy are considered in detail as examples. We discuss two sorts of disorder: i) that in intradimer coupling constants $\cal J$ or in the value of one-ion anisotropy $\cal D$ and ii) disorder in small exchange coupling constants $J_{ij}$ between spins from different dimers or spins on neighboring sites (in large-$\cal D$ systems). For disorder in $\cal J$ or $\cal D$ only, we derive in the first order in the defects concentration $c$ the following expressions for corrections to propagating excitations energies and their damping: Eqs.~\eqref{spec1d} for 1D systems, Eqs.~\eqref{2dcorbot} and \eqref{2dcortop} for 2D systems, and Eqs.~\eqref{spec3dbot} and \eqref{spec3dtop} for 3D ones. It is found that the analytical approach does not work for states near the band edges so that ranges of the analytical results validity are given by Eqs.~\eqref{val1d} in 1D systems and by Eqs.~\eqref{val2d} in 2D and 3D ones. We demonstrate by performing numerical calculations that imaginary parts of the Green's function $\chi''({\bf k},\omega)$ show non-Lorentzian peaks at momenta for which the analytical approach does not work. Analysis of the corresponding wave functions demonstrates the localized nature of states in the band near its edges (see Fig.~\ref{loc2d} for the 2D system). Other states in the band remains propagating in 2D systems (and the same result is expected for 3D ones). In contrast, all states in the band turn out to be localized in 1D bosonic systems that resembles the situation in 1D electronic systems. Besides, we find numerically that the analytical approach does not work in 1D systems if $c|u/aJ|\agt1$ due to multiple-defects resonance scattering that leads to anomalies in corrections to the spectrum and DOS (see Fig.~\ref{dos1}(b)). Analytical consideration of DOS shows that a localized impurity level arises above and below the band for any positive and negative $u$, respectively, in 1D and 2D systems whereas only $|u|>2a|J|$ leads to the isolated level in 3D systems. Taking into account also the disorder in $J_{ij}$, we obtain in 1D systems for the spectrum and the damping Eqs.~\eqref{spec1d22}. Eqs.~\eqref{spec2dbot} and \eqref{spec3dbot2} give the spectrum and the damping in 2D and 3D systems, respectively, near the spectrum minimum, whereas Eqs.~\eqref{spec2dtop} and Eqs.~\eqref{spec3dtop2} are corresponding expressions in the vicinity of the spectrum maximum. In all dimensions, we find a phenomenon of mutual reduction of corrections to the spectrum and the damping from two types of disorder when certain relations are fulfilled involving $u$ and $u_1$. For disorder in $J_{ij}$ only, analytical results for DOS show that one impurity level above the band and one impurity level below it appear if $u_1$ lies outside the interval $-2<u_1/J<0$ in 1D and 2D systems and outside the interval $-2.75<u_1/J<0.75$ in 3D systems. There are no isolated impurity levels for $u_1$ lying inside these intervals. Notice that expressions for the spectrum of propagating modes should also work at small temperature in the vicinity of $H_{bg1}$ or $H_{bg2}$ (see Fig.~\ref{th}). If there are no impurity levels inside the gap, the gap value can be reduced to zero by magnetic field. As a result the ratio of the long-wavelength quasiparticle damping to its energy can reach the value of $c/k^2$ (for 2D systems) in a wide range of parameters. Although this ratio is much smaller than unity in the range of this result validity $1\gg k\gg \sqrt c$ (as it must be for propagating excitations) it is much greater than $c$, the maximum value of $\gamma_{\bf k}/\varepsilon_{\bf k}$ obtained before for long-wavelength magnons in magnetically ordered magnets. \cite{wan,2dvac,syromyat1,syromyat2} The results obtained can be relevant to other gapped phases in bond disordered spin systems both with and without a long range magnetic order. For instance, the phenomenon of localization of states near the band edges was observed theoretically in ferromagnets with random easy-axis anisotropy. \cite{medved} Eqs.~\eqref{spec1d22}, \eqref{spec2dbot}, \eqref{spec2dtop}, \eqref{spec3dbot2}, and \eqref{spec3dtop2} are derived using the general form of the spectrum \eqref{spec0} near its minimum (maximum) and using the general form of the impurity operators \eqref{v1} and \eqref{v2}. Then, they can be used for analysis of other gaped phases in other systems. Results of recent neutron measurements of quasiparticles spectra at $H<H_{c1}$ in bond disordered dimer systems IPA-Cu(Cl$_x$Br$_{1-x}$)$_3$ (Ref.~\cite{nafradi}) and $\rm (C_4H_{12}N_2)Cu_2(Cl$$_{1-x}$Br$_x$)$_6$ (Ref.~\cite{huv2}) were interpreted under assumption that all excitations in the band are conventional wavepackets. As we see above for 1D systems, localized state can behave as a short-wavelength wavepacket. However such a behavior observed experimentally for states lying near the band bottom (corresponding to long-wavelength quasiparticles in pure systems) is quite puzzling. Our results demonstrate pronounced non-Lorentzian shape of Green's function imaginary part for states near the band bottom. Even according to the general theorem \cite{theorem} such states should be localized in these materials because no impurity levels arise in the gap (see Introduction). This point needs further experimental and theoretical analysis. Another point we leave for future studies is the influence of the quasiparticle interaction at low-field phases. This interaction is expected to play important role in real systems in which the gap value at $H=0$ is of the order of the band width. \begin{acknowledgments} This work is supported by RSF grant No.\ 14-22-00281, RFBR Grants No.\ 12-02-01234 and No.\ 12-02-00498. A.V.\ Sizanov acknowledges Saint-Petersburg State University for research grant 11.50.1599.2013. \end{acknowledgments}
\section{Velocities from the variation of observed galaxy luminosities} To linear order in perturbation theory, the observed redshift $z$ of a galaxy typically deviates from its cosmological redshift $z_{c}$ according to (\cite[Sachs \& Wolfe 1967]{SW1967}) \begin{displaymath} \frac{z-z_{c}}{1+z} = \frac{V(t,r)}{c} - \frac{\Phi(t,r)}{c^2} - \frac{2}{c^2}\int_{t(r)}^{t_0}{\rm d}t \frac{\partial\Phi\left\lbrack\hat{\boldsymbol{r}}r(t),t\right\rbrack}{\partial t}\approx \frac{V(t,r)}{c}, \end{displaymath} where $V$ is the (physical) radial peculiar velocity of the galaxy, $r$ is a unit vector along the line of sight to the object, and $\Phi$ denotes the usual gravitational potential. Here we explicitly assume low redshifts such that the velocity $V$ is the dominant contribution, and we further consider all fields relative to their present-day values at $t_{0}$. As the shift $z-z_{c}$ enters the calculation of distance moduli ${\rm DM}=25+5\log_{\rm 10}\lbrack D_{L}/{\rm Mpc}\rbrack$, where $D_{L}$ is the luminosity distance, observed absolute magnitudes $M$ differ from their true values $M^{(t)}$. We thus have \begin{displaymath} M = m - {\rm DM}(z) - K(z) + Q(z) = M^{(t)} + 5\log_{10}\frac{D_{L}(z_{c})}{D_{L}(z)}, \end{displaymath} where $m$ is the apparent magnitude, the function $Q(z)$ accounts for luminosity evolution, and $K(z)$ is the $K$-correction (\cite[Blanton \& Roweis 2007]{Blanton2007}). On scales where linear theory provides an adequate description, the variation $M-M^{(t)}$ of magnitudes distributed over the sky is systematic, and therefore, contains information on the peculiar velocity field. Given a suitable parameterized model $V(\hat{\boldsymbol{r}}, z)$ of the radial velocity field, the idea is now to maximize the probability of observing galaxies with magnitudes $M_{i}$ given only their redshifts and angular positions $\hat{\boldsymbol{r}}_{i}$ on the sky, i.e., \begin{displaymath} P_{\rm tot} = \prod\limits_{i}P\left (M_{i}\vert z_{i}, V(\hat{\boldsymbol{r}}_{i},z_{i})\right ) = \prod\limits_{i}\left (\phi(M_{i})\middle /\int_{M_{i}^{+}}^{M_{i}^{-}}\phi(M){\rm d}M\right ), \end{displaymath} where we assume that redshift errors can be neglected (\cite[Nusser \etal\ 2011]{Nusser2011}), $\phi(M)$ denotes the galaxy luminosity function (LF), and the corresponding limiting magnitudes $M^{\pm}$ depend on $V(\hat{\boldsymbol{r}}, z)$ through the cosmological redshift $z_{c}$. Here the motivation is to obtain a maximum-likelihood estimate of $V(\hat{\boldsymbol{r}}, z)$ by finding the set of velocity model parameters which minimizes the spread in the observed magnitudes. \cite[Tammann \etal\ (1979)]{Tammann1979} first adopted this approach to estimate the motion of Virgo relative to the local group, and recently, \cite[Nusser \etal\ (2011)]{Nusser2011} used it to constrain bulk flows in the local Universe from the 2MASS Redshift Survey (\cite[Huchra \etal\ 2012]{Huchra 2012}). \section{Constraints on the cosmic peculiar velocity field at \boldmath{$z\sim 0.1$}} Galaxies from the Sloan Digital Sky Survey (SDSS) Data Release 7 (\cite[Abazajian \etal\ 2009]{Abazajian2009}) probe the cosmic velocity field out to $z\sim 0.1$. Here we report results obtained from applying the luminosity method to a subset of roughly half a million galaxies (\cite[for additional details, see Feix \etal\ 2014]{Feix2014}). {\underline{\it Data}}. In our analysis, we used the latest version of the NYU Value-Added Galaxy Catalog (NYU-VAGC; \cite[Blanton \etal\ 2005]{Blanton2005}). Giving the largest spectroscopically complete galaxy sample, we adopted (Petrosian) $^{0.1}r$-band magnitudes, and chose the subsample NYU-VAGC {\tt safe} to minimize incompleteness and systematics. Our final sample contained only galaxies with $14.5 < m_{r} < 17.6$, $-22.5 < M_{r} - 5\log_{10}h < -17.0$, and $0.02 < z < 0.22$ (relative to the CMB frame). In addition, we employed a suite of galaxy mock catalogs mimicking the known systematics of the data. {\underline{\it Radial velocity model}}. We considered a bin-averaged velocity model $\tilde{V}(\hat{\boldsymbol{r}})$ in two redshift bins, $0.02 < z < 0.07$ and $0.07 < z < 0.22$. For each bin, the velocity field was further decomposed into spherical harmonics, i.e. \begin{displaymath} a_{lm} = \int{\rm d}\Omega\tilde{V}(\hat{\boldsymbol{r}})Y_{lm}(\hat{\boldsymbol{r}}),\qquad \tilde{V}(\hat{\boldsymbol{r}}) = \sum\limits_{l,m}a_{lm}Y_{lm}^{*}(\hat{\boldsymbol{r}}),\qquad l>0, \end{displaymath} where the sum over $l$ is cut at some maximum value $l_{\rm max}$. Because the SDSS data cover only part of the sky, the inferred $a_{lm}$ are not statistically independent. The impact of the angular mask was studied with the help of suitable galaxy mock catalogs. The monopole term ($l=0$) was not included since it is degenerate with an overall shift of magnitudes. {\underline{\it LF estimators}}. Reliably measuring the galaxy LF represents a key step in our approach. To assess the robustness of our results with respect to different LF models, we analyzed the data using LF estimators based on a Schechter form and a more flexible spline-based model, together with several combinations and variations thereof. For simplicity, we also assumed a linear dependence of the luminosity evolution with redshift. {\underline{\it Bulk flows and higher-order velocity moments}}. Accounting for known systematic errors in the SDSS photometry, our ``bulk flow'' measurements are consistent with a standard $\Lambda$CDM cosmology at a $1$--$2\sigma$ confidence level in both redshift bins. A joint analysis of the corresponding three Cartesian components confirmed this result. To characterize higher-order moments as well, we further obtained direct constraints on the angular velocity power spectrum $C_{l} = \langle |a_{lm}|^{2}\rangle$ up to the octupole contribution. The estimated $C_{l}$ were found compatible to be with the theoretical power spectra of the $\Lambda$CDM cosmology. \begin{figure}[t] \begin{center} \includegraphics[width=0.95\linewidth]{fig2.eps} \caption{Raw estimates of $\sigma_{8}$ obtained from the NYU-VAGC: shown is the derived $\Delta\chi^{2}$ as a function of $\sigma_{8}$ for both redshift bins (left panel) and the first redshift bin with $0.02<z<0.07$ only (right panel), adopting different estimators of the LF (solid, dashed, and dotted lines).} \label{fig1} \end{center} \end{figure} {\underline{\it Constraints on $\sigma_{8}$}}. Assuming a prior on the $C_{l}$ as dictated by the $\Lambda$CDM model with fixed Hubble constant and density parameters, we independently estimated the parameter $\sigma_{8}$ which determines the amplitude of the velocity field. Due to the presence of a dipole-like tilt in the galaxy magnitudes (\cite[Padmanabhan \etal\ 2008]{Padmanabhan2008}), the obtained raw estimates of $\sigma_{8}$ were expected to be biased toward larger values (Fig.\,\ref{fig1}). After correcting for this magnitude tilt with the help of our mocks (Fig.\,\ref{fig2}), we eventually found $\sigma_{8}\approx 1.1\pm 0.4$ for the combination of both redshift bins and $\sigma_{8} \approx 1.0\pm 0.5$ for the low-$z$ bin only, where the low accuracy is due to the limited number of galaxies. This confirms our method's validity in view of future datasets with larger sky coverage and better photometric calibration. \begin{figure}[b] \begin{center} \includegraphics[width=0.95\linewidth]{fig1.eps} \caption{Distribution of $\sigma_{8}$ estimated from mock galaxy catalogs: shown are the recovered histograms (black lines) and respective Gaussian fits with (solid lines) and without (dashed lines) the inclusion of a systematic (randomly oriented) tilt in the galaxy magnitudes, using the information in both redshift bins (left) and the bin with $0.02<z<0.07$ only (right).} \label{fig2} \end{center} \end{figure} \section{Toward constraints on the linear growth rate} A very interesting aspect of our luminosity-based approach is the possibility to place bounds on the growth rate of density perturbations, $\beta = f(\Omega)/b$ (where $b$ is the linear galaxy bias), by modeling the large-scale velocity field directly from the observed clustering of galaxies in redshift space (\cite[Nusser \& Davis 1994]{Nusser1994}). Such bounds are complementary to and --- regarding ongoing and future redshift surveys --- expected to be competitive with those obtained from redshift-space distortions (\cite[Nusser \etal\ 2012]{Nusser2012}). To get an idea of how well the method could constrain $\beta$ at $z\sim 0.1$ from SDSS galaxies, we used mocks generated from the Millennium Simulation (\cite[Springel \etal\ 2005]{Springel2005}; \cite[Henriques \etal\ 2012] {Henriques2012}) to create full-sky catalogs which otherwise shared all characteristics of the real SDSS data. Adopting a radial velocity model proportional to the true one smoothed over spheres of $10h^{-1}$ Mpc radius, the luminosity method was applied to samples with around $2\times 10^{5}$ galaxies and correctly recovered the velocity field. The error on the proportionality constant typically yielded $\pm 0.2$--$0.3$ if only the contribution of multipoles with $l > 25$ (an appropriate value for the SDSS geometry) is taken into account. Assuming an accurate velocity reconstruction for these modes, we expect a similar situation for $\beta$. A further complication is that the angular mask may introduce bias as a consequence of multipole mixing. This and other technical issues mainly related to the reconstruction of the velocity field are currently under detailed investigation. \section{Outlook} Current and next-generation spectroscopic surveys are designed to reduce data-inherent systematics because of larger sky coverage and improved photometric calibration in ground- and space-based experiments (e.g, \cite[Levi \etal\ 2013]{Levi2013}; \cite[Laureijs \etal\ 2011]{Laureijs2011}). The method considered here does not require accurate redshifts and can be used with photometric redshift surveys such as the 2MASS Photometric Redshift catalog (2MPZ; \cite[Bilicki]{Bilicki2014} \cite[\etal\ 2014]{Bilicki2014}) to recover signals on scales larger than the spread of the redshift error. Together with our results, these observational perspectives give us confidence that the luminosity-based method will be established as a standard cosmological probe, independent from and alternative to the more traditional ones based on galaxy clustering, gravitational lensing and redshift-space distortions.
\section*{Introduction} Let $G$ be a simple Lie group and let $G/B$ be the flag variety attached to $G$. These varieties enjoy many nice properties; in particular, they are spherical, i.e.~the Borel subgroup acts on a flag variety with an open orbit. The varieties $G/B$ can be degenerated in such a way that the action of the Borel subgroup degenerates (modulo torus) into an action of the abelian unipotent group of the same dimension, acting with an open orbit on the degenerate flag variety (see\cite{F1}, \cite{F2}). The construction is of Lie theoretic nature and uses the theory of highest weight $G$-modules. It was shown in \cite{CFR1} that in type $A$ the degenerate flag varieties are closely related to the representation theory of the equioriented quivers of type $A$. More precisely, the degeneration of $SL_n/B$ is isomorphic to a certain quiver Grassmannian of subrepresentations of a direct sum of an injective and projective representations. This observation, on the one hand, allows to use the representation theory of quivers in order to study the geometry of the degenerate flag varieties, and, on the other hand, produces links between Lie theory and quiver Grassmannians. Both the degenerate flag varieties and the quiver Grassmannians are known to be related to the theory of Schubert varieties. More precisely, one can identify certain quiver Grassmannians and degenerate flag varieties with certain Schubert varieties (see \cite{CL}, \cite{G}, \cite{PRS}). The main question we ask in this paper is to what extent one can generalize the degeneration constructions above to the case of the affine Lie algebras and affine Lie groups? Is there a connection with the theory of quivers and with Schubert varieties in this case? We give partial answers to the questions above. In short, the affine story is rather complicated (even in type A); the correct replacement of the equioriented type $A$ quiver is the (equioriented) cycle quiver. In our paper we concentrate on the case of the degenerate affine Grassmannians (though the general case is also discussed). From the point of view of representation theory this means that we restrict to the basic level one module of the affine Lie algebras. The general definition of the degenerate flag varieties goes through the theory of highest weight modules. In the affine case the same definition works perfectly. Unfortunately, we are not able to identify the resulting ind-variety even for affine $\msl_n$ (however, we completely describe the $A_1$ case and put forward a conjecture for the symplectic Lie algebras). The problems pop up even on the level of representations: we do not have good enough description of the PBW graded level one basic representation $L_0^a$ of the affine $\msl_n$. In section 2 we formulate a conjecture saying that $L_0^a$ can be realized inside the semi-infinite wedge space. Unfortunately, we can not prove the conjecture at the moment. It turns out that there exists a similar object attached to the affine Lie algebra $\widehat{\mgl}_n$. More precisely, using the formalism of the semi-infinite wedge spaces, we define the PBW degeneration ${\rm Gr}^a(\mgl_n)$ of the affine Grassmannians. The definition is very similar to the general Lie theoretic one and we are able to describe the degenerate object in linear algebra terms. Using this description, we make a connection to the theory of quiver Grassmannians for the one vertex loop quivers. More precisely, ${\rm Gr}^a(\mgl_n)$ can be realized as an inductive limit of quiver Grassmannians, which are analogues of Schubert varieties. We use the representation theoretic techniques to derive algebro-geometric properties of this finitization. In particular, we study the structure of orbits and construct desingularizations explicitly. We also identify them with certain classical affine Schubert varieties for larger groups. Yet another approach to the study of the $\widehat{\mgl}_n$ degenerate Grassmannians comes from the identification with the closure of the semi-infinite orbit for the classical parabolic flag varieties for $\widehat{\msl}_{2n}$. We identify the finitization as above with concrete subvarieties inside the closure. We thus realize ${\rm Gr}^a(\mgl_n)$ as a semi-infinite orbit closure inside the $\msl_{2n}$ classical affine Grassmannian. The paper is organized as follows. In section 1 we collect the basic objects and constructions on affine algebras and PBW degenerations. In section 2 we define and study degenerate affine Grassmannians of type $A$. Section 3 is devoted to the quiver part of the story: the principal quiver Grassmanians for the one vertex loop quivers are studied and the identification with subvarieties in the degenerate flag varieties and with affine Schubert varieties is constructed. In section 4 we discuss the degenerate affine Grassmannians of types $\widehat{\msl}_2$ and $\widehat{\msp}_{2n}$. \section{The setup} \subsection{Affine Lie algebras} Let $\g$ be a simple Lie algebra, $\gh$ the corresponding affine algebra. We have \[ \gh=\g\T\bC[t,t^{-1}]\oplus\bC K\oplus \bC d, \] where $K$ is central and $[d,x\T t^i]=-ix\T t^i$. Let $\g=\fn^-\oplus\fh\oplus\fn$ be the Cartan decomposition. Consider the decomposition for the affine algebra $\widehat{\fg}=\widehat{\fg}^-\oplus\widehat{\fg}^0\oplus\widehat{\fg}^+$, where \begin{gather*} \widehat{\fg}^-=\fn^-\T 1\oplus \fg\T t^{-1}\bC[t^{-1}], \ \widehat{\fg}^0=\fh\T 1\oplus\bC K\oplus \bC d,\\ \widehat{\fg}^+=\fn^+\T 1\oplus \fg\T t\bC[t]. \end{gather*} Let $\theta$ be the highest root for $\fg$. For a dominant integral $\g$-weight $\la$ and a non-negative integer $k$ such that $(\la,\theta)\le k$, let $L_{\la,k}$ be the corresponding irreducible integrable highest weight $\gh$-module with a highest weight vector $v_{\la,k}$. We have \begin{gather*} \widehat{\fg}^+ v_{\la,k}=0,\ \U(\widehat{\fg}^-)v_{\la,k}=L_{\la,k},\\ (h\T 1) v_{\la,k}=\la(h)v_{\la,k},\ Kv_{\la,k}=kv_{\la,k},\ dv_{\la,k}=0. \end{gather*} Let $G$ be a simple simply-connected Lie group with a Borel subgroup $B$, $\fg=Lie(G)$, $\fb=Lie(B)$. Let $\widehat G$ and $I$ be the corresponding affine group and its Iwahori subgroup, respectively. For a parahoric subgroup $P\subset \widehat G$ such that $I\subset P\subset \widehat G$, let $\widehat G/P$ be the corresponding affine flag variety. These varieties are infinite-dimensional ind-varieties, i.e.~they are inductive limits of the finite-dimensional Schubert varieties. More precisely, let $T\subset\widehat G$ be the Cartan torus and let $p$ be a $T$-stable point in the affine flag variety. The corresponding Schubert variety is the closure of the $I$-orbit through $p$. If $P=P_0$ is the maximal parahoric corresponding to the affine simple root then the corresponding flag variety is called the affine Grassmannian. Finally, we note that if $P$ stabilizes the highest weight line $\bC v_{\la,k}$ in $\bP(L_{\la,k})$, then we have a natural embedding $\widehat G/P\subset \bP(L_{\la,k})$. \subsection{Sato Grassmannians and flag varieties of type $A_\infty$} We consider an infinite-dimensional vector space $V$ with a basis $v_i$, $i\in\bZ$. The semi-infinite wedge space $F=\Lambda^{\infty/2}$ is spanned by the elements \[ v_{i_0}\wedge v_{i_1}\wedge\dots,\ i_0>i_1>\dots,\ i_{k+1}=i_k-1 \text{ for } k \text{ large enough.} \] The charge of the wedge product $v_{i_0}\wedge v_{i_1}\wedge\dots$ is defined as $(i_k+k)$ for $k$ large enough. We have the decomposition $F=\bigoplus_{m\in\bZ} F^{(m)}$, where $F^{(m)}$ is spanned by the wedge products of charge $m$. We define the vacuum vectors \[ |m\ket=v_m\wedge v_{m-1}\wedge v_{m-2}\wedge\dots\in F^{(m)}. \] The Lie algebra $\mgl_\infty$ is spanned by the matrix units $E_{i,j}$, i.e. it consists of the infinite matrices with finite support. The natural action of $\mgl_\infty$ on the space $V$ extends to an action on each space $F^{(m)}$. It is easy to see that each $F^{(m)}$ is an irreducible $\mgl_\infty$ module with highest weight vector $|m\ket$. The Sato Grassmannian ${\rm SGr}_m$ sits inside $\bP(F^{(m)})$ via the Pl\"ucker embedding; it consists of subspaces $U\subset V$ with the property that there exists $N\in\bZ$ such that \[ {\rm span}(v_N,v_{N-1},\dots)\subset U \text{ and } \dim U/{\rm span}(v_N,v_{N-1},\dots)=m-N. \] In particular, the line $\bC|m\ket$ belongs to (the image of) ${\rm SGr}_m$. It is easy to see that all the varieties ${\rm SGr}_m$ are isomorphic (via shifts of the indices of $v_k$). Following \cite{KaPe} we define the flag variety $\Fl_\infty$ of type $A_\infty$ as the subvariety of the product $\prod_{m\in\bZ} {\rm SGr}_m$ consisting of collections $(U_m)_{m\in\bZ}$, $U_m\in {\rm SGr}_m$ such that $U_m\subset U_{m+1}$. \subsection{Type $A^{(1)}_{n-1}$ case} Now let us consider an $n$-dimensional vector space $W$ with a basis $w_1,\dots,w_n$. Let us identify the space $W\T \bC[t,t^{-1}]$ with $V$ by \begin {equation}\label{fold} v_{nk+j}=w_j\T t^{-k-1},\ j=1,\dots,n,\ k\in\bZ \end{equation} In particular, \[ |0\ket=(w_1\T 1)\wedge\dots\wedge (w_n\T 1)\wedge (w_1\T t)\wedge\dots \wedge (w_n\T t)\wedge\dots . \] This gives an embedding $\msl_n\T\bC[t,t^{-1}]\subset\mgl_\infty$ and induces an action of $\widehat{\msl}_n$ on each $F^{(m)}$. The irreducible highest weight $\widehat{\msl}_n$ modules can thus be realized in the semi-infinite picture. In particular, the level one module $L_{\omega_i,1}$, $i=0,\dots,n-1$ can be seen as the subspace of $F^{(i)}$ generated by $|i\ket$. We also see that the affine Grassmannian $\widehat{SL_n}/P_0$ of type $A_n$ is naturally embedded into the Sato Grassmannian. The image of this embedding can be described explicitly: \begin{equation} \widehat{SL_n}/P_0 = \{U\in {\rm SGr}_0:\ tU\subset U\}. \end{equation} The complete flag variety for affine $SL_n$ is formed by the collections $(U_i)_{i\in\bZ}$, $U_i\in {\rm SGr}_i$ such that $U_i\subset U_{i+1}$, $tU_i\subset U_{i+n}$. \subsection{PBW filtration and degenerate flag varieties} We first collect some facts on the PBW degeneration in type $A$ (\cite{FFoL1}, \cite{FFoL3}, \cite{F1}). Given a highest weight representation $V_\la$ of $\msl_n$ we have an increasing filtration on $V_\la$, induced by the PBW filtration on $\U(\fn^-)$. The associated graded spaces $V^a_\la$ are modules over the abelian algebra $(\fn^-)^a$ with the underlying vector space $\fn^-$. The degenerate flag variety $\Fl^a_\la(\fg)\subset \bP(V^a_\la)$ is the closure of the orbit of the group $\exp((\fn^-)^a)$ through the line containing the highest weight vector $v_\la$. Let $\la$ be a regular dominant weight for $\fg=\msl_n$. Then the corresponding degenerate flag variety does not depend on $\la$ and can be described explicitly as follows. Let $W=\bC^n$ with a basis $w_1,\dots,w_n$ and let $pr_k$ be the projection along $w_k$ to the span of the remaining basis vectors. Then the degenerate flag variety attached to a regular dominant weight consists of collections $(V_1,\dots,V_{n-1})$, $V_k\in{\rm Gr}_k(W)$ such that $pr_{k+1}V_k\subset V_{k+1}$. We denote this variety by $\Fl^a(\msl_n)$. Note that it was shown recently in \cite{CL}, Theorem 1.2 that $\Fl^a(\msl_n)$ is isomorphic to a Schubert variety for $\msl_{2n-2}$. We also note that the degenerate Grassmann varieties for $\msl_n$ coincide with their classical analogues. The explicit realization of $\Fl^a(\msl_n)$ can be naturally generalized to the $A_\infty$ case. Recall the basis $v_i$ of $V=\bC^\infty$. We define the operators $pr_k:V\to V$ as projections along $v_k$ to the span of $v_i$, $i\ne k$. The following definitions are degenerate versions of the classical analogues (see e.g. \cite{KaPe}). \begin{dfn} The full degenerate flag variety $\Fl^a_\infty$ of type $A_\infty$ is the subvariety of the product $\prod_{m\in\bZ} {\rm SGr}_m$ consisting of collections $(U_m)_{m\in\bZ}$, $U_m\in {\rm SGr}_m$ such that $pr_{m+1}U_m\subset U_{m+1}$. \end{dfn} We note that these varieties can be seen as infinite limits of the type $A_n$ degenerate flag varieties. \begin{rem} The degeneration procedure from $\Fl_\infty$ to $\Fl^a_\infty$ can be described as follows. Let ${\bf F}_\infty$ be the subvariety inside ${\mathbb A}^1\times \prod_{m\in\bZ} {\rm SGr}_m$, consisting of the points $\left(t,(U_m)_{m\in\bZ}\right)$ such that $pr_{m+1}(t) U_m\subset U_{m+1}$, where $pr_{m+1}(t):V\to V$ is the map defined by $pr_{m+1}(t)v_k=v_k$, $k\ne m+1$ and $pr_{m+1}(t)v_{m+1}=tv_{m+1}$. Then we have the natural projection ${\bf F}\to{\mathbb A}^1$. The fiber over $t=0$ is isomorphic to $\Fl^A_\infty$ and the general fiber is isomorphic to $\Fl_\infty$. \end{rem} Let us now define the affine degenerate flag varieties of type $\mgl_n$. We denote by $pr_{w_i\T t^k}:W\T \bC[t,t^{-1}]\to W\T \bC[t,t^{-1}]$ the projections along $w_i\T t^k$ to the span of the remaining vectors of the form $w_j\T t^l$. \begin{dfn} The degenerate flag variety $\Fl^a(\mgl_n)$ of type $\mgl_n$ consists of collections $(U_i)_{i=0}^n$ such that $U_i\in{\rm SGr}_i$, $pr_{w_{i+1}\T t^{-1}} U_i\subset U_{i+1}$ and $U_n=t^{-1}U_0$. \end{dfn} \begin{rem}\label{flgr} We note that if $(U_i)_{i=0}^n\in \Fl^a(\mgl_n)$, then \[ pr_{W\T 1} tU_0 = pr_{w_1\T 1}\dots pr_{w_n\T 1} tU_0\subset U_0. \] Indeed, \[ pr_{W\T t^{-1}} tU_n = pr_{w_1\T t^{-1}}\dots pr_{w_n\T t^{-1}}(U_0)\subset U_n. \] Since $tU_n=U_0$ we obtain $pr_{W\T 1} tU_0\subset U_0$. \end{rem} Now let us consider the case of the affine Lie algebras. The standard PBW filtration $F_\bullet$ on the universal enveloping algebra $\U(\widehat{\fg}^-)$ induces a filtration on the module $L_{\la,k}$. The associated graded space $L^a_{\la,k}$ is a representation of the abelian Lie algebra $\widehat{\fg}^{-,a}$ (with the underlying vector space $\widehat{\fg}^{-}$). \begin{lem}\label{nilp} For any $x\in \widehat{\fg}^{-,a}$ there exists $N$ such that $x^N v_{\la,k}=0$ in $L_{\la,k}^a$. \end{lem} \begin{proof} Recall that the algebra $\widehat{\fg}^{-,a}$ is abelian. It suffices to prove the lemma for $x=r\T t^{-l}$ for some $r\in\fg$, $l>0$. Since $r$ can be included into an $\msl_2$-triple, we may assume $\fg=\msl_2$. Let $e,h,f$ be the standard basis of $\msl_2$. We know that for any $i$ there exists $M$ such that $(f\T t^{-i})^Mv_{\la,k}=0$. Also we have the adjoint action of $\msl_2=\msl_2\T 1$ on the symmetric algebra of $\msl_2$, generating new relations. Hence, it suffices to prove that the algebra Sym$(\msl_2)/(U(\msl_2)f^M)$ is finite-dimensional. But this algebra is known to be finite-dimensional by \cite{F3}, Corollary 4.3. \end{proof} \begin{rem} We note that this lemma does not hold in the non-degenerate situation. \end{rem} Let $\widehat{G}^{-,a}=\exp(\widehat{\fg}^{-,a})$ be the Lie group of the Lie algebra $\widehat{\fg}^{-,a}$. This group is isomorphic to the sum of (an infinite number of) copies of the groups $\bG_a$ (one copy for each negative root of $\gh$). Let us now consider the projective space $\bP(L^a_{\la,k})$. For a vector $v\in L^a_{\la,k}$ let $[v]\in \bP(L^a_{\la,k})$ be the corresponding line in the projectivization. \begin{dfn}\label{dafv} The degenerate affine flag variety $\Fl^a_{\la,k}$ is the closure of the orbit $\widehat{G}^{-,a}[v_{\la,k}]$ inside $\bP(L^a_{\la,k})$. \end{dfn} \begin{rem} The orbit $\widehat{G}^{-,a}[v_{\la,k}]$ makes sense, since the action of the group $\widehat{G}^{-,a}$ on $\bP(L^a_{\la,k})$ is well defined. In fact, thanks to Lemma \ref{nilp}, all the operators from the Lie algebra $\widehat{\fg}^{-,a}$ of the Lie group $\widehat{G}^{-,a}$ act as nilpotent operators and hence can be exponentiated, giving rise to the $\widehat{G}^{-,a}$-action. \end{rem} \begin{rem} We do not describe the degeneration procedure in this paper. Note however that for finite-dimensional Lie algebras of type $A_n$ the closure of the orbit of the group $\widehat{G}^{-,a}$ (of its finite-dimensional analogue, to be precise) is indeed a (flat) degeneration of the classical flag variety (see \cite{F1}, Proposition 5.8). \end{rem} \section{Degenerate affine Grassmannians of type $A$} \subsection{Main theorems} \begin{dfn} The degenerate affine Grassmannian ${\rm Gr}^a(\mgl_n)$ of type $\mgl_n$ is the subvariety of the Sato Grassmannian ${\rm SGr}_0$ consisting of subspaces $U$ such that $pr_{W\T 1} tU\subset U$, where $pr_{W\T 1}$ is the projection along $W\T 1$ to the span of $W\T t^i, i\ne 0$. \end{dfn} We also define finite-dimensional approximations of the varieties ${\rm Gr}^a(\mgl_n)$. \begin{dfn} ${\rm Gr}^a_N(\mgl_n)$ is the subvariety of ${\rm Gr}^a(\mgl_n)$ consisting of $U$ such that \[ W\T t^N\bC[t] \subset U\subset W\T t^{-N}\bC[t]. \] \end{dfn} \begin{rem}\label{B} Let $S_{N,n}=\frac{W\T t^{-N}\bC[t]}{W\T t^N\bC[t]}$. Then ${\rm Gr}_N^a(\mgl_n)$ is isomorphic to the variety of $Nn$-dimensional subspaces $U\subset S_{N,n}$ such that $pr_{W\T 1} tU\subset U$. \end{rem} We note that ${\rm Gr}^a(\mgl_n)$ is naturally the inductive limit of its finite dimensional pieces ${\rm Gr}^a_N(\mgl_n)$. We will prove the following theorems: \begin{thm}\label{fdim} \begin{itemize} \item ${\rm Gr}^a_N(\mgl_n)$ is an irreducible projective variety of dimension $Nn^2$. \item ${\rm Gr}^a_N(\mgl_n)$ carries an action of an abelian unipotent group $\bG_a^{Nn^2}$ with an open dense orbit. \item ${\rm Gr}^a_N(\mgl_n)$ is isomorphic to an affine Schubert variety for the group $\widehat{SL}_{2n}$. \end{itemize} \end{thm} \begin{thm}\label{infdim} \begin{itemize} \item ${\rm Gr}^a(\mgl_n)$ carries an action of an infinite dimensional abelian unipotent group with an open dense orbit. The group is the inductive limit of the groups $\bG_a^{Nn^2}$ from Theorem \ref{fdim}. \item ${\rm Gr}^a(\mgl_n)$ is isomorphic to a semi-infinite orbit closure inside the affine Grassmannian for the group $\widehat{SL}_{2n}$. \end{itemize} \end{thm} \begin{lem}\label{irr} ${\rm Gr}^a_N(\mgl_n)$ is an irreducible projective variety of dimension $Nn^2$. \end{lem} \begin{proof} We prove the claim by constructing a resolution of singularities. We consider the space $S_{N,n}=\frac{W\T t^{-N}\bC[t]}{W\T t^N\bC[t]}$. The operator $pr_{W\T 1} t$ naturally acts on $S_{N,n}$. In particular, $(pr_{W\T 1}t)^N=0$. Let $S_{N,n}(k)={\rm Im} (pr_{W\T 1})^k$ for $k=0,1,\dots,N-1$. In particular, $\dim S_{N,n}(k)=n(2N-2k)$. Let $R_{N,n}$ be the variety of collections of vector spaces $(U_0,U_1,\dots,U_N)$ such that the following holds: \begin{enumerate} \item $U_i\subset S_{N,n}(k),\; \dim U_i=n(N-k)$, \item $pr_{W\T 1}t U_i\subset U_{i+1},\; i=0,\dots, N-1$. \end{enumerate} One easily sees that $R_{N,n}$ is an $N$-floor tower of fibrations over a point, each fibration having fiber $Gr(n,2n)$. In addition, $R_{N,n}$ surjects onto ${\rm Gr}^a_N(\mgl_n)$ (forgetting all $U_i$ with $i>0$) and this surjection is generically one-to-one. In fact, let $R_{N,n}^0\subset R_{N,n}$ be the subvariety cut out by the condition $\dim (pr_{W\T 1})^k t U_0=n(N-k)$. By definition of $R_{N,n}$, the restriction of the surjection $R_{N,n}\to {\rm Gr}^a_N(\mgl_n)$ to $R_{N,n}^0$ is one-to-one. Now easily sees that $R_{N,n}^0$ is an open dense part of $R_{N,n}$. This proves the lemma. \end{proof} \begin{rem} We generalize the claim of Lemma \ref{irr} in Example \ref{XGr} and Corollary \ref{geocon}. \end{rem} In order to prove the remaining statements of Theorem \ref{fdim} and Theorem \ref{infdim} we construct a projective embedding of ${\rm Gr}^a(\mgl_n)$ and hence of its finite-dimensional pieces ${\rm Gr}^a_N(\mgl_n)$. \subsection{Semi-infinite abelianization} Let us decompose the Lie algebra $\mgl_\infty$ into four blocks \[ \mgl_\infty=\mgl_\infty^{-,-}\oplus \mgl_\infty^{-,+}\oplus \mgl_\infty^{+,-}\oplus\mgl_\infty^{+,+}, \] where $\mgl_\infty^{-,+}$ is spanned by the matrix units $E_{i,j}$ with $i\le 0$, $j>0$, $\mgl_\infty^{+,-}$ is spanned by the matrix units $E_{i,j}$ with $i>0$, $j\le 0$ and similarly for the two other summands. For example, $\mgl_\infty^{+,-}$ is spanned by the operators mapping $v_{\le 0}$ to $v_{>0}$. \begin{rem} The summand $\mgl_\infty^{+,-}$ is abelian. \end{rem} Let $p:\mgl_\infty\to \mgl_\infty^{+,-}$ be the projection along the three other summands. Recall the embedding $\mgl_n\T\bC[t,t^{-1}]\subset\mgl_\infty$. Combining this embedding with $p$ we obtain a map \[ \mgl_n\T t^{-1}\bC[t^{-1}]\to \mgl_\infty^{+,-}. \] We define the space \[ L_0(\mgl^a_n)=U(p(\mgl_n\T t^{-1}\bC[t^{-1}]))|0\ket. \] Since $\mgl_\infty^{+,-}$ is abelian, $L_0(\mgl^a_n)$ is a cyclic representation of the abelian Lie algebra $\mgl^a_n\T t^{-1}\bC[t^{-1}]$. \begin{example}\label{h(z)} Let us consider the Lie algebra ${\widehat{\mgl_1}}$. This is nothing but the Heisenberg algebra. Let $F^{(0)}$ be the Fock module. Let us denote by $h_i$, $i\in\bZ$ the basis of ${\widehat{\mgl_1}}$. Then the elements $h_i$, $i<0$ are represented by the formula $h_i=\sum_{k\in\bZ} E_{k-i,k}$. The projection $p$ to $\mgl_\infty^{+,-}$ defines a representation of the abelian Lie algebra spanned by $h_i$, $i<0$ on $F^{(0)}$. This representation is no longer irreducible. In particular, the subspace generated from the highest weight vector is defined by the relation $(h_{-1}+zh_{-2}+\dots)^2=0$. Here $z$ is a variable and the relation above means that the coefficients of $z^k$ vanish for all $k\ge 0$. \end{example} Let us consider the affine Lie algebra $\widehat{\msl}_{2n}$ and its vacuum representation $L_0(\msl_{2n})$ with the highest weight vector $l_0$. Let us embed $\mgl_n^a$ into $\msl_{2n}$ as the unipotent radical corresponding to the weight $\omega_n$. Explicitly, we consider the abelian subalgebra $\fa$ in $\msl_{2n}$, spanned by the matrix units $E_{i,j}$, $n+1\le i\le 2n$, $1\le j\le n$. Clearly, $\fa\simeq \mgl_n^a$. \begin{lem}\label{2n} The identification $\fa\simeq \mgl_n^a$ induces an isomorphism between $L_0(\mgl_n^a)$ and the subspace $U(\fa\T t^{-1}\bC[t^{-1}])l_0$ of $L_0(\msl_{2n})$. \end{lem} \begin{proof} Consider the space $R$ with basis $r_1,\dots,r_{2n}$. We write $R=R^<\oplus R^>$, where $R^<$ is spanned by $r_1,\dots,r_n$ and $R^>$ is spanned by $r_{n+1},\dots,r_{2n}$. Then $L_0(\msl_{2n})$ sits inside $\Lambda^{\infty/2}(R\T\bC[t,t^{-1}])$ with $l_0$ being the wedge product \[ l_0=\bigwedge_{i\ge 0} \left(r_1\T t^i\wedge\dots\wedge r_{2n}\T t^i\right). \] Let us write $R\T\bC[t,t^{-1}]$ as a direct sum of four subspaces $R^{<,+}$, $R^{<,-}$, $R^{>,+}$ and $R^{>,-}$, where \begin{gather*} R^{<,+}=R^<\T \bC[t],\ R^{<,-}=R^<\T t^{-1}\bC[t^{-1}] \end{gather*} and similarly for $R^{>,+}$ and $R^{>,-}$. For example, $l_0$ is the wedge product of the ``top'' wedge powers of $R^{<,+}$ and of $R^{>,+}$. Now let us look at the subspace $U(\fa\T t^{-1}\bC[t^{-1}])l_0$. We take $a\in \fa$ and consider the vector $(a\T t^i) l_0$. Clearly, the only nontrival terms showing up come from the action of $a\T t^i$ sending $R^{<,+}$ to $R^{>,-}$. Recall the space $W$ with the basis $w_1,\dots,w_n$ used to construct the wedge representation for $\widehat{\mgl}_n$. Let us embed $W\T \bC[t,t^{-1}]$ into $R\T\bC[t,t^{-1}]$ as follows: \[ w_j\T t^i\mapsto r_j\T t^i,\ i\ge 0,\ w_j\T t^i\mapsto r_{j+n}\T t^i,\ i< 0. \] Thus the image of this embedding coincides with $R^{<,+}\oplus R^{>,-}$. Now it remains to note that the operators of the form $p(\mgl_n^a)$ map $W\T \bC[t]$ to $W\T t^{-1}\bC[t^{-1}]$. \end{proof} Let us consider the Lie group $GL_n^a(t^{-1}\bC[t^{-1}])=\exp(\mgl_n^a\T t^{-1}\bC[t^{-1}])$. The group $GL_n^a(t^{-1}\bC[t^{-1}])$ is isomorphic to the direct sum of an infinite number of copies of the additive group $\bG_a$ of the base field. Consider the action of $GL_n^a(t^{-1}\bC[t^{-1}])$ on the projectivization $\bP(L_0(\mgl_n^a))$ and (temporarily) denote by $G(n)$ the closure of the orbit through the line containing the highest weight vector. Our goal is to identify $G(n)$ with ${\rm Gr}^a(\mgl_n)$. Let us define a finite-dimensional approximation of $G(n)$ as follows: \[ G_N(n)=\overline{\exp\left(\bigoplus_{i=1}^N \mgl_n^a\T t^{-i}\right)\cdot\bC|0\ket}. \] \begin{lem}\label{GGr} Let $U\in G_N(n)$. Then $pr_{W\T 1}(tU)\subset U$ and $W\T t^N\bC[t] \subset U\subset W\T t^{-N}\bC[t]$. \end{lem} \begin{proof} Take an element $g=\exp(\sum_{i=1}^N x_i\T t^{-i})$, $x_i\in\mgl_n$. Then the space corresponding to the line $g\cdot\bC|0\ket$ is spanned by the vectors \[ w_j\T t^k + x_iw_j\T t^{k-i},\ j=1,\dots,n,\ i=1,\dots,N,\ k=0,\dots,N-1,\ k-i<0. \] Now the claim is clear. \end{proof} \begin{prop}\label{group} $G_N(n)\simeq{\rm Gr}_N^a(\mgl_n)$. \end{prop} \begin{proof} We know that both $G_N(n)$ and ${\rm Gr}_N^a(\mgl_n)$ are irreducible. According to Lemma \ref{GGr}, $G_N(n)$ sits inside ${\rm Gr}^a(\mgl_n)$. Also note that the open orbit \[ \exp\left(\bigoplus_{i=1}^N \mgl_n^a\T t^{-i}\right)\cdot\bC|0\ket \] consists of all $U\in \frac{W\T t^{-N}\bC[t]}{W\T t^N\bC[t]}$ such that the Pl\"ucker coordinate of $U$, corresponding to the set of vectors $w_j\T t^i$, $i\le 0$ does not vanish. Hence the open parts of $G_N(n)$ and ${\rm Gr}^a(\mgl_n)$ coincide. \end{proof} \noindent{\bf Proof of Theorem \ref{fdim}} The first claim follows from Lemma \ref{irr} and the second claim follows from Proposition \ref{group}. The proof of the last claim is postponed until the end of the next section (see Corollary \ref{aSv}). \qed \noindent{\bf Proof of Theorem \ref{infdim}} Since $G_N(n)\simeq{\rm Gr}_N^a(\mgl_n)$, we obtain $G(n)\simeq{\rm Gr}^a(\mgl_n)$. By Lemma \ref{2n} we obtain that ${\rm Gr}^a(\mgl_n)$ is naturally embedded into the classical affine Grassmannian for the group $\widehat{SL}_{2n}$. The image of this embedding is the closure of the orbit of the abelian unipotent group $\exp(\fa\T t^{-1}\bC[t^{-1}])$ through the highest weight line. \qed \subsection{$\msl_n$ case}\label{tilde} Let $L_0(\msl_n)$ be the basic level one $\msl_n$ module with highest weight vector $l_0$. Recall the PBW graded version $L_0^a(\msl_n)$; in particular, $L_0^a(\msl_n)=U(\msl_n^a\T t^{-1}\bC[t^{-1}])l_0$. We define \[ \tilde L_0(\msl_n)=U(p(\msl_n\T t^{-1}\bC[t^{-1}]))|0\ket. \] \begin{conj}\label{T} The $\fn^{-,a}\T t^{-1}\bC[t^{-1}]$-modules $\tilde L_0(\msl_n)$ and $L_0^a(\msl_n)$ are isomorphic. \end{conj} We prove the following lemma. \begin{lem} There exists a surjection $L_0^a(\msl_n)\to \tilde L_0(\msl_n)$ of $\msl_n^a\T t^{-1}\bC[t^{-1}]$ modules. \end{lem} \begin{proof} We need to show that any relation which holds in $L_0^a(\msl_n)$ is true in $\tilde L_0(\msl_n)$. We consider a grading on the semi-infinite wedge space $\Lambda^{\infty/2}(V)$. The $s$-th graded component is spanned by vectors $v_{i_1}\wedge v_{i_2}\wedge\dots$ such that the number of positive indices among $i_l$ is equal to $s$. For example, the zeroth component is spanned by $|0\ket$. Then any $x\in \mgl_\infty^{+,-}$ increases this grading by one. The operators from the three remaining summands either decrease the grading or preserve it. Now assume that we have a relation in $L_0^a(\msl_n)$. To make it explicit, let us fix a basis $x_1,x_2,\dots$ in $\fn^{-}\T t^{-1}\bC[t^{-1}]$. Then any relation can be represented by a homogeneous total degree $N$ polynomial $p(x_1,x_2,\dots)$, which vanishes in $L_0^a$. This means that we have an equality in $L_0$: \[ p(x_1,x_2,\dots) |0\ket = q(x_1,x_2,\dots) |0\ket \text{ for some } q \text{ of degree less than } N. \] This equality implies that $p(x_1,x_2,\dots) |0\ket$ vanishes in $\tilde L_0(\msl_n)$, since this is exactly the degree $N$ component of $p(x_1,x_2,\dots) |0\ket$, considered as a vector in $L_0(\msl_n)$. \end{proof} \begin{rem} To complete the proof of Conjecture \ref{T} it thus suffices to show that the character of $L_0^a(\msl_n)$ is equal to the character of $\tilde{L}_0(\msl_n)$. Unfortunately, we are not able to do this at the moment. \end{rem} \subsection{Flatness} In this section we discuss the infinite-dimensional analogue of the flatness of the degeneration of the $\msl_n$ flag varieties. We consider the space $W\T \bC[t,t^{-1}]$, $W=\bC^n$. Let $U\in{\rm SGr}_0$ be a subspace with $tU\subset U$. We start with the open cell containing the base point $U_0=W\T \bC[t]$. The coordinates in this cell are given by the collection of linear mappings $(A_{k,i})$, $k\ge 1$, $i\ge 0$, $A_{k,i}\in{\rm End} (W)$, such that for a fixed $i$, the operators $A_{k,i}$ vanish for $k$ large enough. The subspace $U$ corresponding to a collection $(A_{k,i})$ is defined as the linear span of the vectors \begin{equation}\label{wti} w\T t^i + A_{1,i}w\T t^{-1}+A_{2,i}w\T t^{-2}+\dots. \end{equation} Now we introduce the operator $\al_\hbar\in ({\rm End} W\T \bC[t,t^{-1}])$ defined by \[ \al_\hbar (w\T t^i)= \begin{cases} w\T t^i,\text{ if } i\ne 0,\\ \hbar w\T t^0,\text{ if } i=0. \end{cases} \] We consider the ind-subscheme ${\bf Gr}\subset{\rm SGr}_0\times\BA^1$ in the product of the Sato Grassmannian and the affine line with coordinate $\hbar$ cut out by the equations $\al_\hbar tU\subset U$. We note that if $\hbar\ne 0$ then the fiber of ${\bf Gr}$ over $\hbar$ is isomorphic to the affine Grassmannian for $\mgl_n$. The special fiber ${\bf Gr}_0$ is the degenerate affine Grassmannian. \begin{prop} \label{flatn} The morphism ${\bf Gr}\to\BA^1$ is flat. \end{prop} \begin{proof} For each connected component ${\bf Gr}^n,\ n\in{\mathbb Z}$, of ${\bf Gr}$, it suffices to produce a dense open affine subscheme $S^n\subset{\bf Gr}^n$ such that $\hbar\in\bC[S^n]$ is not a zero divisor. We will exhibit $S^0$; the other connected components are taken care of similarly. We define $S^0$ as the intersection of ${\bf Gr}^0$ with the open cell in ${\rm SGr}_0$ formed by the subspaces transversal to $W\otimes t^{-1}\bC[t^{-1}]$. \begin{lem} Let $U\in {\rm SGr}_0$ be defined by the collection of operators $(A_{k,i})$. Then $(U,\hbar)\in{\bf Gr}$ if and only if \begin{equation}\label{Aki} A_{k,i}=A_{k-1,i+1}+\hbar A_{k-1,0}A_{1,i} \end{equation} for all $k,i\ge 1$. \end{lem} \begin{proof} We need \begin{equation}\label{tU} w\T t^{i+1} +\hbar A_{1,i}w\T t^{0}+A_{2,i}w\T t^{-1}+\dots\in U. \end{equation} This means that the vector \eqref{tU} is a linear combination of the vectors \eqref{wti} starting with $w\T t^{i+1}$ and $w\otimes t^0$ with coefficients 1 and $\hbar A_{1,i}$. This gives the desired system of equations. \end{proof} \begin{rem} If $\hbar=0$ then equation \eqref{Aki} reduces to $A_{k,i}=A_{k-1,i+1}$ (block Hankel matrices). \end{rem} Now the ind-scheme $S^0$ is a union of the finite-type subschemes $S^0(N)$ cut out by the conditions $A_{k,i}=0$ if $k>N$ or $i\ge N$. The relations \eqref{Aki} allow to express any operator $A_{k,i}$ with $k>0$ in terms of $A_{1,i}$. Hence the coordinate ring $\bC[S^0]$ is a certain completion of the polynomial ring in $\hbar$ and the matrix coefficients of $A_{1,0},A_{1,1},A_{1,2},\ldots$ The relations~\eqref{Aki} are homogeneous if we set $\deg\hbar=0,\ \deg A_{k,i}=k+i$. If we had a relation $\hbar f=0$ for $0\ne f\in\bC[S^0]$, it would imply $\hbar f^{(n)}=0$ for each homogeneous component $f^{(n)}$ of $f$. Since each $f^{(n)}$ is an element of the polynomial ring in $\hbar$ and the matrix coefficients of $A_{1,0},A_{1,1},A_{1,2},\ldots$, we conclude $f^{(n)}=0$, hence $f=0$. The proposition is proved. \end{proof} \begin{rem} The special fiber ${\bf Gr}_0$ is a union of reduced finite type schemes of growing dimensions. For $\hbar\ne0$, in the case of $\mgl_1$, the fiber ${\bf Gr}_\hbar$ is a union of zero-dimensional schemes with nilpotents. So the flatness of ${\bf Gr}$ over ${\mathbb A}^1$ holds only at the level of inductive limit, and looks somewhat counterintuitive. \end{rem} \begin{rem} Again let us consider the case of $\mgl_1$. Then for $\hbar\ne 0$ the fiber lives in the projectivization of the Fock module $\bC[h_{-1},h_{-2},\dots]$ (the Fock module can be naturally identified with the space of sections of the tautological line bundle). However, in the degenerate situation the space of sections of the tautological line bundle is smaller (more precisely, $\bC[h_{-1},h_{-2},\dots]/(h(z)^2)$, i.e. the ideal we quotient out is generated by all the coefficients of the series $(\sum_{i>0} z^{i-1}h_{-i})^2$, see Example \ref{h(z)}). \end{rem} \section{Quiver Grassmannians for (truncated) loop quivers} \subsection{Basics} For $N\geq 1$, let $A_N={\bC}[t]/(t^N)$ be the truncated polynomial ring, which we view as a finite-dimensional and self-injective algebra over ${\bC}$ ($A_N$ being an indecomposable projective and injective module over itself). For a finitely generated (possibly non-commutative) algebra $A$, a finite-dimensional $A$-module $M$ and an integer $k\leq\dim M$, we denote by ${\rm Gr}^A_k(M)$ the Grassmannian of $k$-dimensional subrepresentations $U$ of $M$. This is a projective variety, admitting a closed embedding into the Grassmannian ${\rm Gr}_k(M)$ of $k$-dimensional ${\bC}$-linear subspaces of $M$. Let $W$ be an $m$-dimensional vector space, and consider $W\otimes A_N$ as an $A_N$-module, which is thus projective and injective of dimension $mN$. Our aim is to study the following varieties and to relate them to degenerate affine Grassmannians and to affine Schubert varieties: \begin{definition} With notation as above, define $$X^{(N)}_{k,m}={\rm Gr}_{k}^{A_N}(W\otimes A_N)$$ as the Grassmannian of $k$-dimensional $A_N$-subrepresentations of $W\otimes A_N$. \end{definition} \begin{example}\label{XGr} $X^{(N)}_{Nn,2n}\simeq {\rm Gr}^a_N(\mgl_n)$ (see Remark \ref{B}). \end{example} Let us interpret this definition in linear algebra terms: consider the operator $\varphi$ on $W\otimes A_N$ which is given by $\varphi={\rm id}_W\otimes t$, where $t$ means multiplication by $t$ on $A_N$ (which is a regular nilpotent operator on $A_N$). Thus $\varphi$ is nilpotent and its Jordan canonical form consists of $m$ nilpotent Jordan blocks of size $N$ each. The variety $X^{(N)}_{k,m}$ parametrizes $\varphi$-invariant $k$-dimensional subspaces of $W\otimes A_N$, thus it is naturally a closed subvariety of ${\rm Gr}_k(W\otimes A_N)$.\\ Note that the group ${\rm GL}_m(A_N)\simeq {\rm GL}(W\otimes A_N)$ acts on $X^{(N)}_{k,m}$; this group is of dimension $mN$, with reductive part ${\rm GL}_m({\bC})$ and unipotent radical $1+Mat_{m}(tA_N)$.\\ Our aim in the next subsection is to realize $X^{(N)}_{k,m}$ as a geometric quotient of a well-known variety by a (free) group action. We recall the prototype for such quotient realizations. Let $V$ be a $k$-dimensional vector space. Then the linear Grassmannian ${\rm Gr}_k(W\otimes A_N)$ can be viewed as the quotient of the set ${\rm Hom}^0(V,W\otimes A_N)$ of injective linear maps from $V$ to $W\otimes A_N$ modulo the action of ${\rm GL}(V)$, thus $${\rm Gr}_k(W\otimes A_N)\simeq {\rm Hom}^0(V,W\otimes A_N)/{\rm GL}(V).$$ \subsection{Interpretation as framed moduli} We recall a result of classical invariant theory, see \cite{LBR}, Section 5.1:\\ For two vector spaces $V$ and $W$, consider the action of ${\rm GL}(V)$ on ${\rm End}(V)\times {\rm Hom}(V,W)$ by $g\cdot(\varphi,f)=(g\varphi g^{-1},fg^{-1})$. We call a pair $(\varphi,f)$ stable if $\bigcap_{i\geq 0}{\rm Ker}(f\varphi^i)=0$. This is equivalent to the map $$\bigoplus_{i=0}^{\dim V-1}f\varphi^i:V\rightarrow W^{\dim V}$$ being injective. Denote by $({\rm End}(V)\times {\rm Hom}(V,W))^{\rm st}$ the open subset of stable points; the action of ${\rm GL}(V)$ is free on this stable locus. \begin{thm} The set $({\rm End}(V)\times {\rm Hom}(V,W))^{\rm st}$ admits a geometric quotient by ${\rm GL}(V)$. It embeds into the Grassmannian ${\rm Gr}_{\dim V}(W^{\dim V})$ of subspaces of $W^{\dim V}$ of dimension $\dim V$, by mapping the class of $(\varphi,f)$ to the subspace ${\rm Im}(\bigoplus_{i=0}^{\dim V-1}f\varphi^i)$. \end{thm} Let $\mathcal{N}(V)$ be the closed subvariety of nilpotent operators in ${\rm End}(V)$, and let $\mathcal{N}^{(N)}(V)$ be the set of operators $\varphi$ such that $\varphi^N=0$, for $N\geq 1$. As restrictions of (geometric) quotients to invariant closed subvarieties are again (geometric) quotients, we can consider the ${\rm GL}(V)$-invariant subset $(\mathcal{N}^{(N)}(V)\times{\rm Hom}(V,W))^{\rm st}$ of $({\rm End}(V)\times {\rm Hom}(V,W))^{\rm st}$ and get, after some reindexing and using $N$-nilpotency, the following result: \begin{cor} The set $(\mathcal{N}^{(N)}(V)\times {\rm Hom}(V,W))^{\rm st}$ admits a geometric quotient by ${\rm GL}(V)$. This quotient embeds into the Grassmannian ${\rm Gr}_{\dim V}(W\otimes A_N)$ of $dim V$-dimensional subspaces of $W[[t]]/(t^N)\simeq W\otimes A_N$ by mapping the class of $(\varphi,f)$ to the subspace ${\rm Im}(\sum_{i=0}^{N-1}f\varphi^{N-1-i}t^i)$. \end{cor} The main technical result of this section is the following (compare with \cite{L}, section 2): \begin{lem} The image of the above embedding coincides with the subvariety $X^{(N)}_{\dim V,\dim W}$ of ${\rm Gr}_{\dim V}(W\otimes A_N)$. \end{lem} \begin{proof} For $\varphi$ and $f$ such that $\varphi^N=0$, the image of $\sum_if\varphi^{N-1-i}t^i$ is $t$-invariant, namely $$t\sum_{i=0}^{N-1}f\varphi^{N-1-i}t^i=\sum_{i=0}^{N-1}f\varphi^{N-1-i}t^{i+1}=\sum_{i=0}^{N-1}f\varphi^{N-i}t^i=(\sum_{i=0}^{N-1}f\varphi^{N-1-i}t^i)\varphi,$$ and thus $$t\cdot{\rm Im}(\sum_if\varphi^{N-1-i}t^i)={\rm Im}(t\sum_if\varphi^{N-1-i}t^i)=$$ $$={\rm Im}((\sum_i f\varphi^{N-1-i}t^i)\varphi)\subset{\rm Im}(\sum_if\varphi^{N-1-i}t^i).$$ Conversely, let the image of an injective map $\sum_{i=0}^{N-1}f_it^i$ be $t$-invariant. This means that there exists an endomorphism $\varphi$ such that $$\sum_{i=1}^{N}f_{i-1}t^i=t\sum_{i=0}^{N-1}f_it^i=(\sum_{i=0}^{N-1}f_it^i)\varphi=\sum_{i=0}^{N-1}f_i\varphi t^i.$$ Comparing coefficients, this is equivalent to $$f_0\varphi=0,\; f_0=f_1\varphi,\, f_1=f_2\varphi,\,\ldots\, f_{N-2}=f_{N-1}\varphi,$$ thus $f_i=f_{N-1}\varphi^{N-1-i}$ for all $i=0,\ldots,N-1$, and $f_{N-1}\varphi^N=0$. The latter conditions means that ${\rm Im}(\varphi^{i+1})\subset{\rm Ker}(f_{N-1}\varphi^{N-1-i})$ for all $i$, thus $${\rm Im}(\varphi^N)\subset \bigcap_{i=0}^{N-1}{\rm Ker}(f_{N-1}\varphi^{N-1-i})=\bigcap_{i=0}^{N-1}{\rm Ker}(f_i)=0$$ by injectivity of $\sum_if_it^i$, which means $\varphi^N=0$. Thus, the pair $(\varphi,f_{N-1})$ maps to the image of $\sum_if_it^i$.\end{proof} We have thus proved: \begin{cor}\label{maincor} The variety $X^{(N)}_{\dim V,\dim W}$ is isomorphic to the quotient $$(\mathcal{N}^{(N)}(V)\times{\rm Hom}(V,W))^{\rm st}/{\rm GL}(V).$$ \end{cor} \subsection{Geometric consequences} The above corollary allows us to easily derive various geometric properties of the varieties $X^{(N)}_{k,m}$: \begin{cor} \label{geocon} The variety $X^{(N)}_{k,m}$ is irreducible, normal, Cohen-Macaulay with rational singularities. It has dimension $\dim\mathcal{N}^{(N)}({\bC}^k)+k(m-k)$. In particular, it has dimension $k(k-1)+k(m-k)=k(m-1)$ for $N\geq k$. \end{cor} \begin{proof} Every variety $\mathcal{N}^{(N)}({\bC}^k)$ is irreducible, normal, Cohen-Macaulay with rational singularities by \cite{KP}, Theorem 0.1, since it is the closure of a single conjugacy class. These properties are preserved under passing to open subsets and geometric ${\rm GL}_k({\bC})$-quotients. Namely, for irreducibility this is clear, whereas normality and rational singularities are preserved under arbitrary quotients (the latter by Boutot's theorem (\cite{B}, Corollaire)). For the Cohen-Macaulay property, we use the fact that a geometric ${\rm GL}_k({\bC})$-quotient is a principal bundle under a special group, and thus Zariski locally trivial (see \cite{Se}, Theoreme 2). Finally, the dimension formula follows by a direct calculation.\end{proof} The natural sequence of embeddings of the varieties $\mathcal{N}^{(N)}(V)$, stabilizing in $\mathcal{N}^{(\dim V)}(V)=\mathcal{N}(V)$, induces a chain of embeddings $$X_{k,m}^{(1)}\subset X_{k,m}^{(2)}\subset \ldots\subset X_{k,m}^{(k)}=X_{k,m}^{(k+1)}=\ldots,$$ whose limit we define as $X_{k,m}=X_{k,m}^{(N)}$ for $N\geq k$, which is the quotient of stable pairs $(\varphi,f)$ for $\varphi$ an arbitrary nilpotent operator by the action of ${\rm GL}_k({\bC})$. \subsection{Examples} We give some examples of the varieties $X_{k,m}$. First, it is easy to see that $X_{1,m}\simeq{\bf P}^{m-1}$. Second, let us consider the case $k=2$. By the above, we thus consider the set of $2m\times 2$-matrices of rank $2$ of the form $$\left[\begin{array}{l}AB\\ A\end{array}\right]$$ for an $m\times 2$-matrix $A$ and a nilpotent $2\times 2$-matrix $B$, up to the ${\rm GL}_2$-action on columns. We can embed this variety into projective space via the Pl\"ucker embedding. Namely, we choose homogeneous coordinates $x_{i,j}$ for $1\leq i<j\leq 2m$ and map the above matrix to the collection of its $2\times 2$-minors $T_{i,j}$. A priori these obey the Pl\"ucker relations $$T_{i,k}T_{j,l}=T_{i,j}T_{k,l}+T_{i,l}T_{j,k}$$ for all $1\leq i<j<k<l\leq 2m$. Since $\det(B)=0$, we have $T_{i,j}=0$ for $1\leq i<j\leq m$. Since ${\rm tr}(B)=0$, we have $T_{i,j+m}=T_{j,i+m}$ for all $1\leq i,j\leq m$. We conclude that $X_{2,m}$ can be realized as the set of points in projective space with coordinates $x_{i,j}$ for $1\leq i<j\leq 2m$ subject to the relations $$x_{i,k}x_{j,l}=x_{i,j}x_{k,l}+x_{i,l}x_{j,k}\mbox{ for }1\leq i<j<k<l\leq 2m,$$ $$x_{i,j}=0\mbox{ for }1\leq i<j\leq m,$$ $$x_{i,j+m}=x_{j,i+m}\mbox{ for }1\leq i,j\leq m.$$ For example, in case $m=2$ we can eliminate the variables $x_{1,2}$ and $x_{2,3}$ and realize $X_{2,2}$ as the singular surface in ${\bf P}^3$ with coordinates $x_{1,3},x_{1,4},x_{2,4},x_{3,4}$ and defining equation $x_{1,3}x_{2,4}=x_{1,4}^2$ (then $(0:0:0:1)$ is an isolated singularity).\\ For general $m$, we can eliminate the variables $x_{i,j}$ for $1\leq i<j\leq m$ and $x_{j,i+m}$ for $1\leq i<j\leq m$. We can rename the remaining variables as $v_{i,j}=x_{i+m,j+m}$ for $1\leq i<j\leq m$ and $w_{i,j}=x_{i,j+m}$ for $1\leq i,j\leq m$ and rewrite the above relations in these terms. Computer experiments with $m\leq 5$ suggest the following: \begin{conj} The variety $X_{2,m}$ is isomorphic to the closed subvariety of projective space with coordinates $v_{i,j}$ for $1\leq i<j\leq m$ and $w_{i,j}$ for $1\leq i\leq j\leq m$ given by the following equations: \begin{enumerate} \item the symmetric matrix $W=(w_{i,j})$ has rank one, \item $WZ=0$ for the matrix $Z$ with $m$ rows, with columns indexed by tuples $(j,k,l)$ for $1\leq j<k<l\leq m$, and with entries $$Z_{i,(j,k,l)}=\left\{\begin{array}{lll}v_{k,l}&,&i=j,\\ -v_{j,l}&,&i=k,\\ v_{j,k}&,&i=l,\\ 0&,&i\not=j,k,l\end{array}\right.$$ \item the Pl\"ucker relations for the $v_{i,j}$. \end{enumerate} \end{conj} \subsection{Orbit structure} Now we consider the action of ${\rm GL}_m(A_N)$ on $X_{k,m}^{(N)}$ and determine the orbit structure. The group ${\rm GL}_m(A_N)$ embeds into the group ${\rm GL}(W\otimes A_N)$ as the subgroup of automorphisms commuting with ${\rm id}_W\otimes t$; every such automorphism can be written uniquely as $$\sum_{i=0}^{N-1}\psi_it^i$$ for $\psi_0\in{\rm GL}(W)$, $\psi_i\in{\rm End}(W)$ for $i=1,\ldots,N-1$, where each summand acts on $W\otimes A_N$ by applying the endomorphism $\psi_i$ in the $W$-component and multiplying by $t^i$ in the $A_N$-component. In particular, such an element acts on a linear map $\sum_if_it^i$ from $V$ to $W\otimes A_N$ by $$(\sum_i\psi_it^i)(\sum_if_it^i)=\sum_i(\sum_{i'+i''=i}\psi_{i'}f_{i''})t^i.$$ This allows us to conclude: \begin{lem}\label{orb} Under the isomorphism $$X^{(N)}_{\dim V,\dim W}\simeq(\mathcal{N}^{(N)}(V)\times{\rm Hom}(V,W))^{\rm st}/{\rm GL}(V),$$ the action of ${\rm GL}_{\dim W}(A_N)$ on $X_{\dim V,\dim W}^{(N)}$ translates to $$(\sum_i\psi_it^i)\cdot\overline{(\varphi,f)}=\overline{(\varphi,\sum_i\psi_if\varphi^i)}$$ for $\sum_i\psi_it^i\in{\rm GL}_{\dim W}(A_N)$, $\varphi\in\mathcal{N}^{(N)}(V)$ and $f\in{\rm Hom}(V,W)$. \end{lem} \begin{proof} Compute the action of $\sum_i\psi_it^i$ on $\sum_if\varphi^{N-1-i}t^i$ as above and compare the $t^{N-1}$-coefficients.\end{proof} To parametrize the orbits of ${\rm GL}_m(A_N)$ in $X_{k,m}^{(N)}$, we use representation theory of the algebra $A_N$ and some of the methods developed in \cite{CFR1}. More precisely, we will use the following facts: \begin{itemize} \item The indecomposable representations of $A_N$ are (up to isomorphism) the $U_i={\bC}[t]/(t^i)$ for $i=1,\ldots,N$. The representation $U_i$ has dimension $i$ and socle $U_1$. We have $\dim{\rm Hom}_{A_N}(U_i,U_j)=\min(i,j)$. In particular, $A_N$ admits only finitely many isomorphism classes of representations of fixed dimension. \item Two subrepresentations $R,R'$ of an injective representation $I$ of a finite-dimensional algebra $A$ are conjugate under ${\rm Aut}(I)$ if and only if they are isomorphic. \item A representation $R$ embeds into an injective representation $I$ of a finite-dimensional algebra $A$ if and only if the socles embed, that is, ${\rm soc}(R)$ embeds into ${\rm soc}(I)$. \item If an $A$-representation $M$ admits only finitely many isomorphism classes of subrepresentations of a given dimension $k$, the Grassmannian of subrepresentations ${\rm Gr}_k^A(M)$ is stratified into locally closed subsets $\mathcal{S}_{[R]}$ consisting of subrepresentations in a fixed isomorphism class $[R]$; we have $\dim\mathcal{S}_{[R]}=\dim{\rm Hom}(R,M)-\dim{\rm End}(R)$. \end{itemize} Combining the above statements, we have: \begin{prop} Suppose $A$ is an algebra of finite representation type, and let $I$ be an injective representation of $A$. Then the ${\rm Aut}(I)$-orbits $\mathcal{O}_{[R]}$ in ${\rm Gr}_k(I)$ are parametrized by the isomorphism classes $[R]$ such that $R$ is $k$-dimensional and ${\rm soc}(R)\subset{\rm soc}(I)$. The orbit $\mathcal{O}_{[R]}$ has dimension $\dim{\rm Hom}(R,I)-\dim{\rm End}(R)$. \end{prop} This proposition applies to our setting since $A=A_N$ is of finite representation type and $I=U_N^m$ is injective. Every $U_i$ has simple socle, thus a representation embeds into $U_N^m$ if and only if it has at most $m$ indecomposable direct summands. A $k$-dimensional such representation can be written as $U_\lambda=U_{\lambda_1}\oplus\ldots\oplus U_{\lambda_m}$ for a partition $N\geq\lambda_1\geq\ldots\geq\lambda_m\geq 0$ with $\sum_i\lambda_i=k$ (note the parts are allowed to be zero).\\ \begin{cor} The ${\rm GL}_m(A_N)$-orbits $\mathcal{O}_\lambda$ in $X_{k,m}^{(N)}$ are parametrized by partitions $\lambda$ of $k$ of length $m$ with parts at most $N$. We have $$\dim\mathcal{O}_\lambda=mk-\sum_i(\lambda_i')^2,$$ where $\lambda'$ denotes the conjugate partition of $\lambda$. \end{cor} \begin{proof} We know that the dimension of $\dim\mathcal{O}_\lambda$ is equal to $$\dim{\rm Hom}(U_\lambda,U_N^m)-\dim{\rm End}(U_\lambda).$$ We note that $\dim{\rm Hom}(U_{\lambda_i},U_N)=\lambda_i$ and $\dim{\rm Hom}(U_{\lambda_i},U_{\lambda_j})=\lambda_{{\rm max}(i,j)}$. Hence $\dim\mathcal{O}_\lambda=(m+1)k-2\sum_{i=1}^m i\lambda_i$. Now using \cite[I.(1,6)]{Mac}, we arrive at the desired formula. \end{proof} Let us determine the closure relation of these orbits. We recall some facts on orderings on partitions from e.g. \cite{Mac}: \begin{itemize} \item We write $\lambda\geq\mu$ (the so-called dominance ordering on partitions) if $\sum_{j\leq i}\lambda_j\geq\sum_{j\leq i}\mu_j$ for all $i$. \item We have $\lambda>\mu$ minimally if and only if there exist entries $i<j$ such that $\mu_i=\lambda_i-1$, $\mu_j=\lambda_j+1$ and $\mu_k=\lambda_k$ for all $k\not=i,j$. \item We have $\lambda\geq\mu$ if and only if for the conjugate partitions one has $\mu'\geq\lambda'$. \end{itemize} Consider partitions $\lambda$, $\mu$ as in the corollary. \begin{thm}\label{adherence} The closure of $\mathcal{O}_\lambda$ contains $\mathcal{O}_\mu$ if and only if $\lambda\geq\mu$. \end{thm} \begin{proof} Suppose $\lambda\geq\mu$. To prove that $\mathcal{O}_\mu$ is contained in the closure of $\mathcal{O}_\lambda$, it suffices to do this in the case where $\lambda\geq\mu$ minimally, thus $\mu$ differs from $\lambda$ only in two positions $i<j$ as above. We can then reduce to the case $m=2$, thus we want to prove that the closure of the orbit corresponding to the partition $(\lambda_1,\lambda_2)$ contains the orbit corresponding to $(\lambda_1-1,\lambda_2+1)$ (in particular, $\lambda_1\ge\lambda_2+2$). We consider the following family $U_z$ for $z\in{\bC}$ of $t$-invariant subspaces of $A_N^2$: the subspace $U_z$ is generated by $(zt^{N-\lambda_1},t^{N-\lambda_1+1})$ and $(t^{N-\lambda_2-1},zt^{N-\lambda_2})$. Then $U_z$ belongs to $\mathcal{O}_{(\lambda_1,\lambda_2)}$ for $z\not=0,1,-1$ and to $\mathcal{O}_{(\lambda_1-1,\lambda_2+1)}$ for $z=0$. (The $z=0$ case is clear; if $z\ne 0$, then $U_z\simeq U_{\lambda_1}\oplus U_{\lambda_2}$ if and only if the vectors $t^{\lambda_1-\lambda_2-1}(zt^{N-\lambda_1},t^{N-\lambda_1+1})$ and $(t^{N-\lambda_2-1},zt^{N-\lambda_2})$ are linearly independent, which is equivalent to $z\ne \pm 1$).\\ To prove the converse, we define for a sequence $k_*=(k_1,\ldots,k_N)$ the subset $$C(k_*)=\{U\in X_{k,m}^{(N)}\, :\, \dim t^iU\leq k_{i+1}\mbox{ for } i=0,\ldots,N-1\}.$$ This is a closed subset since the dimension inequalities can be interpreted as rank conditions. For $U\in\mathcal{O}_\lambda$, we have $$\dim t^iU=\sum_{j:\lambda_j\geq i}(\lambda_j-i).$$ Thus, defining $k_*(\lambda)$ by $$k_i(\lambda)=\sum_{j:\lambda_j\geq i-1}(\lambda_j-i+1)=\sum_{j\geq i}\lambda_j',$$ we have $\mathcal{O}_\lambda\subset C(k_*(\lambda))$. This immediately implies $\overline{\mathcal{O}_\lambda}\subset C(k_*(\lambda))$. Since we already know that $\overline{\mathcal{O}_\lambda}$ contains all $\mathcal{O}_\mu$ for $\lambda\geq\mu$, the claim follows once we prove that $C(k_*(\lambda))$ is contained in the union of the $\mathcal{O}_\mu$ for $\lambda\geq\mu$. So let us assume that $\mathcal{O}_\mu$ is contained in $C(k_*(\lambda))$. Then we know, by the definitions, that $$\sum_{j:\mu_j\geq i}(\mu_j-i)\leq\sum_{j:\lambda_j\geq i}(\lambda_j-i)$$ for all $i=0,\ldots,N-1$; this can be rewritten as $\mu'\geq\lambda'$, thus $\lambda\geq\mu$. The theorem is proved.\end{proof} In particular, we can determine the open orbit of ${\rm GL}_m(A_N)$ in $X_{k,m}^{(N)}$; it corresponds to the partition $(\underbrace{N,\ldots,N}_{s},r,0,\ldots,0)$, where $k=sN+r$ for $0\leq r<N$. \subsection{Desingularization}\label{resolution} For $N,k,m$ as before and a sequence $(k_1,\ldots,k_N)$ as in the proof of Theorem \ref{adherence} (in particular, $k_1=k$), we consider the variety $Y(k_*)$ of tuples $$(U_1,\ldots,U_{N})\in\prod_{i=0}^{N-1}{\rm Gr}_{k_{i+1}}(W\otimes t^iA_N)$$ such that \begin{enumerate} \item $U_1\supset U_2\supset\ldots\supset U_{N}$, \item $t(U_i)\subset U_{i+1}$ for $i=1,\ldots,N-1$ \end{enumerate} (as before, we abbreviate the map ${\rm id}_W\otimes t$ simply by $t$). This is a closed subvariety of the product $\prod_{i=0}^{N-1}{\rm Gr}_{k_{i+1}}(W\otimes t^iA_N)$ and thus projective. \begin{rem} We can view the variety $Y(k_*)$ as a quiver Grassmannian as follows: consider the quiver $\Gamma_N$ with $N$ vertices $v_1,\ldots,v_{N}$ and $2(N-1)$ arrows $\alpha_i:v_i\rightarrow v_{i+1}$ for $i=1,\ldots,N-1$ and $\beta_i:v_{i+1}\rightarrow v_i$ for $i=1,\ldots,N-1$. We consider the admissible ideal $I$ in ${\bC}\Gamma_N$ generated by the elements $$\alpha_i\beta_i-\beta_{i+1}\alpha_{i+1},\, i=1,\ldots,N-1$$ (where for $i=N-1$, the relation has to be read as $\alpha_{N-1}\beta_{N-1}=0$, i.e. we formally define $\alpha_{N}$ and $\beta_{N}$ as zero). The algebra ${\bC}\Gamma_N/I$ is then isomorphic to the Auslander algebra of $A_N$. We construct a specific representation ${\bf W}$ of ${\bC}\Gamma_N/I$: we define ${\bf W}_{v_i}=W\otimes t^iA_N$, ${\bf W}_{\alpha_i}=t$ (multiplication by $t$ in the second component) and ${\bf W}_{\beta_i}=\iota$, the inclusion map. We see that this representation of $\Gamma_N$ indeed satisfies the relations in $I$. Thus we can interpret $Y(k_*)$ as ${\rm Gr}_{k_*}^{{\bC}\Gamma_N/I}({\bf W})$. \end{rem} \begin{prop} For every partition $\lambda$ as above, projection to ${\rm Gr}_{k}(W\otimes A_N)$ induces a desingularization map $$\pi_\lambda:Y(k_*(\lambda))\rightarrow\overline{\mathcal{O}_\lambda}.$$ \end{prop} \begin{proof} We first verify that the image of $Y(k_*)$ under the projection $$\pi:\prod_i{\rm Gr}_{k_{i+1}}(W\otimes t^iA_N)\rightarrow{\rm Gr}_{k}(W\otimes A_N)$$ given by $(U_1,\ldots,U_{N})\mapsto U_1$ is contained in $X_{k,m}^{(N)}$, and even in $C(k_*)$: indeed, the defining relations of $Y(k_*)$ imply that $$t(U_1)\subset U_2\subset U_1,$$ which shows that $\pi$ projects to $X_{k,m}^{(N)}$. More generally, for all $i=0,\ldots,N-1$, iterating the defining relations shows that $$t^{i}(U_1)\subset U_{i+1},$$ which is thus a subspace of dimension at most $k_{i+1}$. But these are precisely the defining conditions of $C(k_*)$.\\ Applying this argument to the special case $k_*=k_*(\lambda)$ for a partition $\lambda$ as before, we see that the image $\pi(Y(k_*(\lambda)))$ is contained in $C(k_*(\lambda))$, which by the proof of Theorem \ref{adherence} equals the closure of the orbit $\mathcal{O}_\lambda$.\\ Again by the proof of Theorem \ref{adherence}, we already know that, for a point $U\in\mathcal{O}_\lambda$, we have $\dim t^i(U)=k_{i+1}(\lambda)$ for $i=0,\ldots,N-1$. Thus, the fibre of $\pi$ over such a point $U$ consists of the single point $(U,tU,\ldots,t^{N-1}U)$ for dimension reasons. First this proves that $\mathcal{O}_\lambda$ is contained in the image of the map $\pi:Y(k_*(\lambda))\rightarrow\overline{\mathcal{O}_\lambda}$. But since $\pi$ is proper, thus has closed image, even $\overline{\mathcal{O}_\lambda}$ is contained in the image, thus (by what is already proven) $\pi$ maps onto $\overline{\mathcal{O}_\lambda}$. Second, the above argument proves that $\pi$ is generically one-to-one.\\ Finally, we prove that $Y(k_*)$ is always smooth, by realizing it as a tower of Grassmann bundles. To do this, we consider truncated versions of the variety $Y(k_*)$, namely, for $i=1,\ldots,N$, we consider the subvariety $Y_i(k_*)$ of $\prod_{j=i-1}^{N-1}{\rm Gr}_{k_{j+1}}(W\otimes t^jA_N)$ given by the same conditions as before, i.e. the subvariety consisting of tuples $(U_i,\ldots,U_{N})$ such that $tU_j\subset U_{j+1}$ for all relevant $j$ and $U_i\supset\ldots\supset U_N$. We have a sequence of projections $$Y(k_*)=Y_1(k_*)\rightarrow Y_2(k_*)\rightarrow\ldots\rightarrow Y_N(k_*).$$ Obviously, we have $$Y_N(k_*)\simeq{\rm Gr}_{k_N}(\underbrace{W\otimes t^{N-1}A_N)}_{\simeq W}).$$ Now we consider the projection $Y_{N-1}(k_*)\rightarrow Y_N(k_*)$, which is equivariant for the action of ${\rm GL}(W\otimes t^{N-1}A_N)$, the latter being transitive on the target. Its fibre over a point $U_N$ consists of all $U_{N-1}$ such that $$U_{N}\subset U_{N-1}\subset t^{-1}U_{N}.$$ Note that $U_{N}\subset t^{-1}U_{N}$ since $tU_N=0\subset U_N$. Thus the fibre over $U_N$ is isomorphic to $${\rm Gr}_{k_{N-1}-k_N}(\underbrace{t^{-1}U_{N}/U_N}_{\simeq W}).$$ We continue inductively: the fibre of a projection $Y_{i-1}(k_*)\rightarrow Y_{i}(k_*)$ over a point $U_{i}$ consists of all $U_{i-1}$ such that $U_{i}\subset U_{i-1}\subset t^{-1}U_{1}$; we note that $tU_{i}\subset U_{i+1}\subset U_i$ and thus $U_i\subset t^{-1}U_{i}$. Thus the fibre over $U_i$ is isomorphic to $${\rm Gr}_{k_{i-1}-k_i}(\underbrace{t^{-1}U_i/U_i}_{\simeq W}).$$ Thus we see that $Y(k_*)$ is a tower of Grassmann bundles with fibres isomorphic to ${\rm Gr}_{k_{i}-k_{i+1}}(W)$ for $i=1,\ldots,N$ (formally defining $k_{N+1}=0$), and in particular, $Y(k_*)$ is irreducible, smooth, of dimension $$mk_1-\sum_i(k_i-k_{i+1})^2.$$ In the special case $k_*=k_*(\lambda)$, this dimension can be written as $mk-\sum_i(\lambda_i')^2$ as expected. \end{proof} \subsection{All $X_{k,m}^{(N)}$ are affine Schubert varieties} Let $P_i$, $i=0,\dots,m$ be the maximal parahoric subgroup of $\widehat{SL}_m$, corresponding to the simple root $\al_i$ of the affine algebra $\widehat{\msl}_m$. We prove the following proposition. \begin{prop}\label{Sch} The variety $X_{k,m}^{(N)}$ is isomorphic to a Schubert variety in $\widehat{SL}_m/P_{k\!\!\!\mod\! m}$. \end{prop} \begin{proof} Recall the Iwahori group $I\subset \widehat{SL}_m$. Consider the $T$-fixed point $x_{k,m}^{(N)}\in X_{k,m}^{(N)}$ such that $Ix_{k,m}^{(N)}$ is the open part of $X_{k,m}^{(N)}$. Namely, fixing the standard basis $e_1,\dots,e_m$ of $\bC^m$ let us write $k=Nr+s$, $0\le j<N$. Then $x_{k,m}^{(N)}$ is the subspace of $\bC^m\T\bC[t]/t^N$ spanned by the vectors $e_i\T t^j$, $i=1,\dots,r-1$, $j=0,\dots,N-1$ and $e_r\T t^j$, $j=N-s,\dots,N-1$. Our goal is to prove the existence of a $T$-fixed point $y_{k,m}^{(N)}\in \widehat{SL}_m/P_{k\!\!\!\mod\! m}$ such that the closure of $Iy_{k,m}^{(N)}$ (the corresponding Schubert variety) is isomorphic to the closure of $Ix_{k,m}^{(N)}$. To this end we construct an embedding of the $I$-module $\Lambda^k(\bC^m\T \bC[t]/t^N)$ into the semi-infinite wedge space with the image contained in the level one representation $L_{k\!\!\!\mod\! m}$ corresponding to the $\msl_m$ fundamental weight $\omega_{k\!\!\!\mod\! m}$. Let us write $k=ma+b$, where $b=k\!\!\!\mod\! m$. Then $X_{k,m}^{(N)}$ contains the span of \[ \bC^m \T t^j, j=t^{N-1},\dots,t^{N-a},\ e_1\T t^{N-a-1},\dots, e_b\T t^{N-a-1}. \] This point is obviously $I$-fixed. We want to identify it with the highest weight line in $L_b$. This line inside $F^{(b)}$ is spanned by $\bC^m \T t^j$, $j\ge 0$ and $w_1\T t^{-1},\dots, w_b\T t^{-1}$. Hence the map $e_i\T t^j\mapsto w_i\T t^{j-N+a}$ induces the $I$-modules embedding of $\Lambda^k(\bC^m\T \bC[t]/t^N)$ to $L_b$, sending $X_{k,m}^{(N)}$ to the Schubert variety $\overline{Iy_{k,m}^{(N)}}\subset \widehat{SL}_m/P_b$. \end{proof} \begin{cor}\label{aSv} ${\rm Gr}_N^a(\mgl_n)$ is an affine Schubert variety for $\widehat{SL}_{2n}$. \end{cor} \begin{rem} We note that Proposition \ref{Sch} implies that $X_{k,m}^{(N)}$ is irreducible, normal, Cohen-Macaulay with rational singularities (see~\cite[Th\'eor\`emes~2.$\Sigma$,~3]{Mat}, \cite[Theorems~2.16,~2.23]{Kum}). In particular, this reproves the first part of Corollary \ref{geocon}. \end{rem} \section{Affine Lie algebras $\widehat{\msl}_2$ and $\widehat{\msp}_{2n}$} \subsection{Type $A_1$} In this subsection we restrict to the case $\g=\msl_2$. We prove Conjecture \ref{T} and give an explicit realization of the degenerate affine Grassmannian inside the Sato Grassmannian $\rm{SGr}_0$. Recall the identification $V\simeq W\T \bC[t,t^{-1}]$, $\dim W=2$. Let $\rm{pr}$ be the projection operator along $W\T 1$ to the span of the basis vectors $w_i\T t^j$, $j\ne 1$. We will also need a skew-symmetric form on $V$ defined by \[ (w_1\T t^i,w_2\T t^j)=\delta_{i+j,-1}, \ (w_1\T t^i,w_1\T t^j)=(w_2\T t^i,w_2\T t^j)=0. \] \begin{thm}\label{sl2} The degenerate affine Grassmannian $\Gr^a(\msl_2)$ sits inside $\rm{SGr}_0$ as the subvariety of subspaces $U$ satisfying the following conditions \begin{enumerate} \item $\rm{pr} (tU)\subset U$, \item $U$ is isotropic with respect to the above symplectic form. \end{enumerate} \end{thm} We first show the existence of the embedding $\Gr^a(\msl_2)\subset \rm{SGr}_0$ by proving Conjecture \ref{T}. To do this we prove that there exists a basis of $L_0(\msl_2)$ such that its vectors are linearly independent when considered in $\tilde L_0(\msl_2)$ (see Section \ref{tilde}). Let $e,h,f$ be the standard basis of $\msl_2$. For an element $x\in\msl_2$ we set $x_i=x\T t^i$. Recall the construction of the $ehf$-basis of $L_0$ from \cite{FKLMM}, Theorem 4. A monomial of the form \begin{equation} \label{form} \dots f_{-n}^{a_n} h_{-n}^{b_n} e_{-n}^{c_n}\dots f_{-1}^{a_1} h_{-1}^{b_1} e_{-1}^{c_1} \end{equation} is called a $ehf$-monomial if it satisfies the following conditions: \begin{enumerate} \item[(a)] $a_i+a_{i+1}+b_{i+1}\le 1$ for $i> 0$, \label{a} \item[(b)] $a_i+b_{i+1}+c_{i+1}\le 1$ for $i> 0$, \label{b} \item[(c)] $a_i+b_i+c_{i+1}\le 1$ for $i> 0$, \label{c} \item[(d)] $b_i+c_i+c_{i+1}\le 1$ for $i> 0$. \label{d} \end{enumerate} Then applying the $ehf$-monomials to a highest weight vector of $L_0$ one gets a basis. The following picture from \cite{FKLMM} illustrates the set of $ehf$-monomials: \begin{center} \begin{picture}(180,70) \multiput(40,20)(0,40){2}{\line(1,0){122}} \multiput(40,20)(20,0){7}{\line(0,1){40}} \multiput(40,20)(20,0){6}{\line(1,1){20}} \multiput(40,20)(20,0){6}{\line(1,2){20}} \multiput(40,40)(20,0){6}{\line(1,1){20}} \multiput(165,20)(0,20){3}{\dots} \multiput(40,20)(20,0){7}{\circle*{3}} \multiput(40,40)(20,0){7}{\circle*{3}} \multiput(40,60)(20,0){7}{\circle*{3}} \put(32,10){$a_1$} \put(52,10){$a_2\ \dots$} \put(32,65){$c_1$} \put(52,65){$c_2 \ \dots$} \end{picture} \end{center} Namely one considers the set of monomials \eqref{form} such that the sum of exponents over the ends of any segment is less than or equal to $1$. \begin{lem} The monomials \eqref{form} subject to the conditions (a)--(d) form a basis of $\tilde L_0$. \end{lem} \begin{proof} Recall the identification $V\simeq W\T \bC[t,t^{-1}]$ given by $v_{2i+1}\mapsto w_1\T t^{-i-1}$, $v_{2i}\mapsto w_2\T t^{-i}$ (not to be confused with $x_k=x\T t^k$ for $x\in\msl_2$). We note that \begin{gather*} h_k v_{2i+1} = v_{2i+1+2k},\ h_k v_{2i} = - v_{2i+2k}, \\ e_k v_{2i+1} = 0,\ e_k v_{2i}= v_{2i+2k-1},\\ f_k v_{2i+1} = v_{2i+2k+1},\ f_k v_{2i}= 0. \end{gather*} Now assume that we are given a monomial $m$ of the form \eqref{form} subject to the conditions (a)--(d). Then $m|0\ket$ is decomposed as a sum of several semi-infinite wedge products of vectors $v_i$. We attach to $m$ one wedge product $w(m)$ from this decomposition. We then show that, given a linear combination of $ehf$-monomials, we can find a monomial $m$ in it such that $w(m)$ does not show up in any other $ehf$-monomial. We note that $f_{-1}$ shifts an index by $1$, $h_{-1}$ by $2$, $e_{-1}$ by $3$, $f_{-2}$ by $3$, $h_{-2}$ by $4$, $e_{-2}$ by $5$ and so on. Now we can see that if $m$ is a $ehf$-monomial, then all powers $a_i$, $b_i$ and $c_i$ are zeroes or ones and moreover, if $x_i$ and $y_j$ show up in $m$, then the difference of their shifts is at least two. Hence $m|0\ket$ contains a semi-infinite wedge product of the form \[ E_{i_1,j_1}\dots E_{i_k,j_k} (v_0\wedge v_1\wedge v_2\wedge\dots),\ j_k>j_{k-1}>\dots >j_1>i_1>\dots>i_k, \] where $E_{i,j}$ are matrix units. Now it is easy to see that given a linear combination of $ehf$-monomials we can find a wedge product as above, contained in a single $ehf$-monomial. \end{proof} Now let us consider the degenerate affine Grassmannian $\Gr^a(\msl_2)$; since affine Grassmannians are special cases of affine flag varieties, the general Definition \ref{dafv} applies (with $k=1$ and $\la=0$). \begin{cor} The $\widehat{\msl}_2$ degenerate Grassmannian is contained in the Sato Grassmannian $\rm{SGr}_0$. \end{cor} \begin{lem}\label{lc} Let $U\in \Gr^a(\msl_2)$. Then $U$ satisfies all the conditions of Theorem \ref{sl2}. \end{lem} \begin{proof} We use the standard basis $w_1,w_2$ of the two-dimensional vector representation of $\msl_2$. In particular, $fw_1=w_2$, $ew_2=w_1$, $hw_1=w_1$, $hw_2=-w_2$. Consider an element $g=\exp(\sum_{i<0} x_ie_{-i}+y_ih_{-i}+z_i f_{-i})$, $g\in \widehat{G}^{-,a}$. Then $g|0\ket$ is spanned by vectors of the form \begin{gather*} w_1\T 1 + y_1w_1\T t^{-1} + z_1w_2\T t^{-1} + y_2w_1\T t^{-2}+z_2w_2\T t^{-2} +\dots,\\ w_2\T 1 + x_1w_1\T t^{-1} - y_1w_2\T t^{-1} + x_2w_1\T t^{-2} - y_2w_2\T t^{-2} +\dots,\\ w_1\T t + y_2w_1\T t^{-1} + z_2w_2\T t^{-1} + y_3w_1\T t^{-2}+z_3w_2\T t^{-2} +\dots,\\ w_2\T t + x_2w_1\T t^{-1} - y_2w_2\T t^{-1} + x_3w_1\T t^{-2} - y_3w_2\T t^{-2} +\dots. \end{gather*} Now its easy to check that the linear span of these vectors satisfies all the conditions of Theorem \ref{sl2}. Since $\Gr(\msl_2)$ is the closure of the $\widehat{G}^{-,a}$ orbit, the lemma follows. \end{proof} \begin{cor} Theorem \ref{sl2} holds. \end{cor} \begin{proof} One sees from the proof of Lemma \ref{lc} that the $\widehat{G}^{-,a}$ orbit of the line $\bC|0\ket$ consists of the subspaces $U$ satisfying the conditions of Theorem \ref{sl2} and such that $U$ has a nontrivial Pl\"ucker coordinate corresponding to the subspace $W\T\bC[t]$. Since this is an open condition, Theorem \ref{sl2} follows. \end{proof} Finally, let us compute the torus acting on the degenerate affine Grassmannian. Assume that we have a torus scaling the basis vectors as $w_1\T t^i\to q_iw_1\T t^i$, $w_2\T t^j\to p_jw_2\T t^j$ for some numbers $p_i$, $q_j$. We want this torus to act on the open cell. In particular, each vector in the list of vectors from the proof of Lemma \ref{lc} has to be invariant (up to scaling) with respect to the torus action. This gives the following set of relations, labeled by positive numbers $k$: \begin{gather*} \frac{p_0}{q_{-k}}=\frac{p_1}{q_{-k+1}}=\dots = \frac{p_{k-1}}{q_{-1}},\\ \frac{q_0}{q_{-k}}=\frac{q_1}{q_{-k+1}}=\dots = \frac{q_{k-1}}{q_{-1}}=\frac{p_0}{p_{-k}}=\frac{p_1}{p_{-k+1}}=\dots = \frac{p_{k-1}}{p_{-1}},\\ \frac{q_0}{p_{-k}}=\frac{q_1}{p_{-k+1}}=\dots = \frac{q_{k-1}}{p_{-1}}. \end{gather*} \begin{rem} The values $p_k=p_1r^{k-1}$, $q_k=q_1r^{k-1}$ for arbitrary $p_1,q_1,r$ solve the equations above. This is a torus (effectively two-dimensional, since it contains one-dimensional torus $r=1$, $p_1=q_1$, scaling all the vectors by the same number), generated by the loop rotation and the Cartan torus of $SL_2$. \end{rem} The system above is equivalent to the statement that for any $a\ge 0$, $b<0$ the quantities $\frac{p_a}{q_b}$, $\frac{q_a}{p_b}$, $\frac{p_a}{p_b}$, $\frac{q_a}{q_b}$ depend only on the difference $a-b$ and $\frac{p_a}{p_b}=\frac{q_a}{q_b}$. This means that there exists a complex number $r$ such that \[ q_a=q_0r^a, p_a=p_0r^a, a\ge 0;\quad q_b=q_{-1}r^{b+1}, p_b=p_{-1}r^{b+1}, b<0 \] with additional condition $\frac{q_0}{q_{-1}}=\frac{p_0}{p_{-1}}$. \begin{cor} The torus acting on the degenerate Grassmannian is 3-dimensional. It is generated by the loop rotation, the one-dimensional Cartan torus of $SL_2$ and an additional one-dimensional torus with $p_{\ge 0}=q_{\ge 0}=1$, $p_{<0}=q_{<0}=\mathrm{const}.$ \end{cor} \subsection{Symplectic degenerate affine Grassmannian} It turns out that the construction of ${\rm Gr}^a(\msl_2)$ has a natural (though conjectural) generalization to the case of symplectic algebras. Let $W$ be a $2n$-dimensional vector space endowed with a non-degenerate skew-symmetric form $(\cdot,\cdot)$. We fix a basis $w_1,\dots,w_{2m}$ of $W$ such that $(w_i,w_{2n+1-i})=1$ for $i=1,\dots,n$. Consider the space $W\T \bC[t,t^{-1}]$ and the corresponding sector $F^{(0)}$ of the semi-infinite wedge power. Now consider the Lie algebra $\msp^a_{2n}$, which is an abelian Lie algebra with the underlying vector space $\msp_{2n}$. Define the action of $\msp_{2n}^a\T t^{-1}\bC[t^{-1}]$ on $W\T \bC[t,t^{-1}]$ as follows. For $x\in \msp_{2n}^a$, $w\in W$ we define \[ (x\T t^i)(w\T t^j)= \begin{cases} xw\T t^{i+j}, \text{ if } j\ge 0, i+j<0,\\ 0,\text{ otherwise}. \end{cases} \] Now let $L_0(\msp_{2n})$ be the basic level one module of the affine Lie algebra $\widehat{\msp}_{2n}$ and let $L_0^a$ be its degenerate analogue. Let $l_0\in L_0$ be the highest weight vector. In particular, $L_0(\msp_{2n})=\U(\msp_{2n}\T t^{-1}\bC[t^{-1}]) l_0$. \begin{conj}\label{sc} Let $\tilde L_0(\msp_{2n})=\U(\msp_{2n}^a\T t^{-1}\bC[t^{-1}])|0\rangle$. Then we have an isomorphism of $\msp_{2n}^a\T t^{-1}\bC[t^{-1}]$-modules \[ \tilde L_0(\msp_{2n})\simeq L^a_0,\ |0\rangle\mapsto l_0. \] \end{conj} The action of the abelian Lie algebra $\msp_{2n}^a\T t^{-1}\bC[t^{-1}]$ as above induces an action of the corresponding infinite-dimensional Lie group $\exp(\msp_{2n}^a\T t^{-1}\bC[t^{-1}])$. Because of Conjecture \ref{sc} the natural candidate for the degenerate symplectic affine Grassmannian ${\rm Gr}^a(\widehat{\msp}_{2n})$ is the closure of the orbit of this group through the highest weight vector inside $\bP(\tilde L_0(\msp_{2n}))$. We denote this closure by ${\rm G}^a(\widehat{\msp}_{2n})$. Define a symplectic form on the infinite-dimensional space $W\T \bC[t,t^{-1}]$: $\bra v\T t^i,w\T t^j\ket =(v,w)\delta_{i+j,-1}$. \begin{thm}\label{sp} The variety ${\rm G}^a(\widehat{\msp}_{2n})$ consists of points $U$ of the Sato Grassmanian $SGr_0$ such that $U$ is isotropic with respect to the form $\bra\cdot ,\cdot\ket$ and $pr(tU)\subset U$. \end{thm} \begin{conj} ${\rm Gr}^a(\widehat{\msp}_{2n})\simeq {\rm G}^a(\widehat{\msp}_{2n})$. \end{conj} We sketch the proof of Theorem \ref{sp}. Let $\rm{O}\subset SGr_0$ be the subvariety of spaces that intersect trivially with $W\T t^{-1}\bC[t^{-1}]$ (i.e. the Pl\"ucker coordinate corresponding to $W\T\bC[t]$ does not vanish). \begin{prop} The $\exp(\msp_{2n}\T t^{-1}\bC[t^{-1}])\cdot \bC |0\ket$ orbit coincides with $\rm{O}\cap \overline{{\rm G}^a(\widehat{\msp}_{2n})}$. \end{prop} In order to prove Theorem \ref{sp} we consider the finitization of ${\rm G}^a(\widehat{\msp}_{2n})$, thus making explicit the ind-variety structure. \begin{dfn} For $N\ge 0$, let ${\rm G}_N^a(\widehat{\msp}_{2n})$ be the finite-dimensional subvariety of ${\rm G}^a(\widehat{\msp}_{2n})$ consisting of subspaces $U$ such that \begin{enumerate} \item $W\T t^N\bC[t]\subset U\subset W\T t^{-N}\bC[t]$, \item $U$ is isotropic, \item $pr(tU)\subset U$. \end{enumerate} \end{dfn} \begin{lem} ${\rm G}_N^a(\widehat{\msp}_{2n})$ coincides with the closure of the orbit of the group $\exp(\msp_{2n}^a\T {\rm span}(t^{-1},\dots,t^{-N}))$ through the line spanned by $|0\ket$. In particular, ${\rm G}_N^a(\widehat{\msp}_{2n})$ are irreducible and of dimension $N\dim\msp_{2n}$. \end{lem} \begin{lem} ${\rm G}_N^a(\widehat{\msp}_{2n})\cap O = \exp(\msp_{2n}^a\T {\rm span}(t^{-1},\dots,t^{-N}))\cdot\bC |0\ket\cap O$. \end{lem} Now it remains to prove the irreducibility of the varieties ${\rm G}_N^a(\widehat{\msp}_{2n})$. This is achieved by constructing explicitly the desingularization via the same procedure as in Lemma \ref{irr} and section \ref{resolution} (see also \cite{FFiL}, Definition 5.1). \section*{Acknowledgments} Thanks are due to L.~Positselski for his explanations about the notion of flatness. We are very grateful to B.~Feigin for extremely fruitful discussions of degenerate affine Grassmannians. The work of E.F. was partially supported by the Dynasty Foundation and by the Simons foundation. The work of EF was supported within the framework of a subsidy granted to the HSE by the Government of the Russian Federation for the implementation of the Global Competitiveness Program. The research of M.F. was carried out at the IITP RAS at the expense of the Russian Foundation for Sciences (project no. 14-50-00150).
\section{Introduction} \parskip=5pt \normalsize The null energy condition (NEC) lies at the origin of the standard picture of early universe's cosmological evolution, determining many of its fundamental properties. For a universe dominated by a perfect fluid, satisfying the NEC is equivalent to the positivity of the sum of energy and pressure $\rho+p>0$, leading to ever-increasing energy density as the evolution is run backward in time. The regime of an $\mathcal{O}(1)$ sensitivity to the short-distance completion of gravitational interactions in the past is thus unavoidable for any NEC-satisfying cosmology. Usually, violating the NEC is synonymous with instabilities -- at least for a system consisting of an arbitrary number of scalar fields with up to one derivative per field in the action \cite{Dubovsky:2005xd,Hsu:2004vr}. The theorem is not without loopholes, though. One possibility of evading it is provided by the \textit{ghost condensate} \cite{ArkaniHamed:2003uy}, that crucially relies on (spontaneously) broken Lorentz invariance in a way that gives rise to a non-standard $\omega \sim k^2$ infrared dispersion relation for the scalar driving the NEC violation. And indeed, it was argued in Ref. \cite{Creminelli:2006xe} that ghost condensation can lead to consistent alternative cosmologies with a weak ($\dot H\ll H^2$) violation of the null energy condition. Another loophole has emerged with the discovery of higher-derivative, yet ghost free scalar theory - the \textit{galileon} \cite{Nicolis:2008in}. The simplest such theory with a cubic self-interaction arises \cite{Luty:2003vm} in the context of the DGP model \cite{Dvali:2000hr}, while the full set of galileons have been found to describe the helicity-0 polarization of the graviton in dRGT theories of ghost-free massive gravity \cite{deRham:2010ik, deRham:2010kj}. It has immediately been realized that (conformal) galileons can be implemented in building a NEC-violating alternative scenario to inflation, referred to as \textit{galilean genesis} (GG) \cite{Creminelli:2010ba}. In this class of models, conformal transformations (or, sometimes, just the dilatations \cite{Creminelli:2012my}) are assumed to be a symmetry of the flat-space theory, nonlinearly realized on a scalar field $\pi$, while couplings of $\pi$ to gravity are assumed to weakly break that symmetry. A crucial difference from inflation is that gravity is largely irrelevant for the early universe, described by GG: the cosmological phase of interest (during which the perturbations relevant for the CMB are produced) effectively takes place on a quasi-Minkowski spacetime, while scale-invariant density perturbations are naturally produced due to the unbroken dilatation invariance of the (time-dependent) scalar background\footnote{Similar ideas lie behind other constructions, such as that of a complex scalar rolling down a negative quartic potential \cite{Rubakov:2009np,Osipov:2010ee}, the \textit{pseudo-conformal universe} \cite{Hinterbichler:2011qk} and DBI genesis \cite{Hinterbichler:2012fr,Hinterbichler:2012yn}. The corresponding NEC-violating backgrounds are characterized by the same symmetry-breaking pattern, albeit technically realized in different ways.}. Moreover, flatness, homogeneity and horizon problems are automatically solved due to the quasi-Minkowski nature of the background spacetime and the gradual shrinking of the comoving Hubble horizon $(a H)^{-1}$. It is thus fair to say that, as far as the standard problems of the Big-Bang cosmology as well as density perturbations are concerned, galilean genesis is degenerate in its predictions with inflation. The differences come with the inclusion of tensor modes: irrelevance of gravity in genesis cosmologies results in a strongly blue-tilted and a completely unobservable (at least as far as the CMB experiments are concerned) spectrum of tensor perturbations \cite{Creminelli:2010ba}. For that reason, it is commonly believed that any possible detection of primordial gravitational waves (such as the one recently claimed by the BICEP2 collaboration \cite{Ade:2014xna}) would strongly disfavor genesis models, as well as their many variations. Indeed, a detectable, scale-invariant tensor spectrum requires that the background spacetime be (quasi-) de Sitter (dS) at the time of freezeout of the relevant set of modes (see, \textit{e.g.} \cite{Creminelli:2014wna} for a recent discussion). In the case that the interpretation of detected $B$-modes as a primordial signal persists, this would mean that any scenario that aims at describing the early universe should allow for a sufficiently extended period of de Sitter evolution. This apparently singles out the standard slow-roll inflation as the preferred paradigm for providing the flat and homogeneous universe with the particle horizon way beyond the observable patch. One motivation of the present work is to re-assess the latter observation, with a focus on galilean genesis as an alternative to inflation. We will broadly define genesis as a phase of the universe with a strongly NEC-violating ($\varepsilon\equiv \dot H/H^2 \geq 1$) expansion that starts out in a low-curvature, maximally symmetric (essentially Minkowski or de Sitter) spacetime. Can such initial conditions result in a scale-invariant and unsuppressed tensor spectrum in a sufficiently broad range of physical scales? As noted above (at least for scalar-tensor theories we will be discussing below) generating scale-invariant tensor modes requires the geometry to be close to de Sitter for a certain period of time during the system's evolution. The question therefore reduces to that of the possibility for the universe to consistently evolve from a low/zero-curvature background in the far past to a much higher curvature inflationary dS spacetime capable of generating observable tensor spectrum at intermediate stages of its history. Because the system has to pass through a quasi de Sitter regime, one should be able to keep good theoretical control over the dynamics beyond the point when gravity starts playing a non-negligible role. Indeed, in the original GG, the moment of time $t_0$ at which gravity becomes order-one important is roughly the moment of the effective field theory (EFT) breakdown and not too long after that the universe is assumed to reheat, while all relevant cosmological perturbations are generated at times $t\ll t_0$ (we will assume time to flow from $t=-\infty$ towards $t=0$ throughout). This situation is sketched by the red curve on Fig. \ref{fig:1}. In terms of the model parameters, \begin{eqnarray} \label{tzero} t_0\sim -\frac{f}{M_{\rm Pl}} \frac{1}{H_0}\, \end{eqnarray} where $f$ is the decay constant of $\pi$, while $H_0\ll f$ is a free parameter, setting the scale for the expansion rate around $t \sim t_0$ (the natural value for the decay constant is $f\sim M_{\rm Pl}$, which we will assume for definiteness in this section). The `slow-roll' parameter $\varepsilon$, starting out formally infinite at $t=-\infty$, decreases with time and is naively estimated to be of order unity at $t_0$. This means that the geometry can not be approximated by de Sitter space at any time during the genesis phase. \begin{figure} \includegraphics[width=0.7\textwidth]{sketch.pdf}\centering \caption{A sketch of the early universe's expansion rate as a function of time for the standard slow-roll inflation (black), as well as original (red) and extended (blue) genesis scenarios.} \label{fig:1} \end{figure} While most of the qualitative features of GG directly follow from scale invariance of the (flat-space) $\pi$-lagrangian, the latter symmetry is badly broken by gravity around $t=t_0$. The background field value can be estimated at that time as \begin{eqnarray} \phi\equiv e^\pi \simeq \mathcal{O}(1)~, \end{eqnarray} whereas throughout the genesis phase $\phi\ll 1$. One is then led to conclude that the loop-generated symmetry-breaking terms in the effective action for $\pi$ itself can start influencing the dynamics for $t\sim t_0$ -- even in the extreme case that these are down by the Planck scale. Indeed, the canonically normalized field $\pi_c$ becomes of order $\pi_c(t_0) \sim f $, making \textit{e.g.} the Planck-suppressed operator $\pi_c (\partial\pi_c)^2$ of the same order as the kinetic term. These estimates motivate extending the $\pi$ action by dilatation-breaking operators that, while irrelevant throughout the genesis phase, could in principle strongly influence the dynamics around the time when gravity becomes order-one important. We will show below that at least for a well-defined subclass of the resulting extensions, cosmological solutions do exist that, while resembling galilean genesis at early times, smoothly extend \textit{beyond} the time $t=t_0$ as illustrated by the blue curve on Fig. \ref{fig:1}. These solutions asymptote, starting from some time $t_{i}$, to an inflationary (quasi) de Sitter space on which both the scalar and the tensor modes are generated with scale-invariant spectrum, just like in inflation. Nevertheless, the scenario at hand -- referred to as \textit{extended genesis} (EG) below -- crucially differs from inflation in that the universe's evolution at early times ($t\ll t_i$) looks nothing like that of the standard NEC-satisfying slow-roll models. Most importantly, NEC violation provides a possibility to avoid the singularity in the past, with the universe gradually relaxing to a low- (or even zero-) curvature space as it is run backwards in time. Due to the latter property, extended genesis can be alternatively viewed as a `UV' (or, to be more precise, as an early-time) -complete realization of inflation. In the cases we consider below, the late-time dynamics of EG will be described by NEC-violating versions of \textit{galileon inflation} (also referred to as \textit{G-inflation}) \cite{Kobayashi:2010cm} -- a model that possesses a number of phenomenologically attractive properties. First, it can produce a large tensor-to-scalar ratio without trans-Planckian field excursions, unlike the standard slow-roll inflation \cite{Lyth:1996im} (because of the shift symmetry, the inflaton itself is not an observable in galileon inflation). Second, similar to ghost \cite{ArkaniHamed:2003uz, Senatore:2004rj} and DBI \cite{Alishahiha:2004eh} models, galileon inflation can lead to a sizeable equilateral nongaussianity. Finally, since $\pi$ itself acquires a scale-invariant spectrum, it is in principle unnecessary to invoke spectator fields (required in many alternatives to inflation) for generating the observed density perturbations. The paper is organized as follows. We start in Sec.\ref{sec2} by spelling out general criteria that a theory, capable of describing the genesis -- de Sitter transition of Fig. \ref{fig:1}, should satisfy. In the same section we give a simple example of a solution with the given feature. Sections \ref{ggal} and \ref{analytic} deal with an analytic construction of such theories, providing explicit examples of completely stable cosmological solutions exhibiting extended genesis. In Sec.\ref{hdim} we study possible effects of higher derivative operators in the effective theory on the scalar spectrum of the backgrounds under consideration. Finally, in Sec.\ref{concl} we conclude. Technical details, that would overwhelm the main body of the text, are collected in the two appendices. The theories described in the rest of the paper are only intended as a starting point for constructing realistic early universe cosmologies based on EG. While we do touch on this in what follows, a fully realistic model-building is left for future work. Most importantly, however, our examples serve as a proof of principle of the possibility to smoothly and stably connect the inflationary quasi-de Sitter universe to a low or even zero-curvature, maximally symmetric spacetime in the asymptotic past. \section{Generalities} \label{sec2} Before diving into a more detailed discussion, we briefly highlight the major properties of theories allowing for the genesis - dS transition. We expect these properties to be the defining ingredient of any other construction capable of achieving our goals. Most importantly, the theories of interest enjoy an enhanced symmetry both for small as well as for large values of the 'sigma model' field $\phi$. In both limits $e^\pi\ll 1$ and $e^\pi\gg 1$, the (flat-space) $\pi$-lagrangian will acquire invariance either under dilatations \begin{eqnarray} \label{dil} \pi(x)\to \pi(e^\lambda x)+\lambda~, \end{eqnarray} describing the scale-invariant (and, in special cases, conformal) galileon \cite{Nicolis:2008in}, or under constant \textit{shifts} \begin{eqnarray} \label{shift} \pi(x)\to \pi(x)+\lambda~, \end{eqnarray} describing $P(X)$ or ordinary galileon-type theories\footnote{By `$P(X)$ theories' we mean theories, defined by their lagrangian being an arbitrary function $P$ of the combination $X\equiv -(\partial\pi)^2$. In the inflationary context these were first studied in \cite{ArmendarizPicon:1999rj}. } with ghost condensation, see \textit{e.g.} \cite{ArkaniHamed:2003uy,ArkaniHamed:2003uz}. Apart from the two (asymptotically) exact symmetries, for $e^\pi\mathrel{\rlap{\lower4pt\hbox{\hskip1pt$\sim$} 1$ the theories under consideration will be \textit{approximately} invariant under internal galilean transformations, \begin{eqnarray} \label{galinv} \pi\to\pi+b_\mu x^\mu~, \end{eqnarray} with $b_\mu$ a constant four-vector. The reason it is useful to think of galilean invariance as an approximate symmetry is that the operator that breaks it has a parametrically suppressed Wilson coefficient in the effective theory. Approximate invariance under \eqref{galinv} then makes this suppression stable under loop corrections, see the discussion below. Galilean invariance becomes more and more pronounced as $\pi\to 0$. As we will see in Sec.\ref{analytic}, in certain cases the small-field regime will itself consist of two qualitatively different stages -- the system gradually evolving from ghost condensate (described by an effectively shift-symmetric theory) in the asymptotic past, into galilean genesis with an enhanced scale invariance \eqref{dil} -- all while $\phi\ll 1$. \begin{comment} \begin{figure} \centering \includegraphics[width=.7\textwidth]{invariances} \caption{Illustration of emergent symmetries on backgrounds with various expectation values of the scalar $\pi$.} \label{symmetries} \end{figure} \end{comment} The asymptotically emergent symmetries are precisely what makes the existence of NEC-violating cosmologies, interpolating between Minkowski and de Sitter spacetimes possible. Let us \textit{e.g.} consider the genesis-de Sitter transition of Fig. \ref{fig:1}. The enhanced conformal invariance at early times/small field values\footnote{To avoid confusion, we note again that `small field values' refers to the expectation value of the sigma model field $e^\pi$, while the goldstone $\pi$ is characterized by large negative values in the given regime. } \textit{generically} gives rise to galilean genesis-like evolution of the universe, whereby conformal invariance, $SO(4,2)$, gets broken down to the maximal de Sitter subgroup $SO(4,1)$ by a time-dependent $\pi$-background\footnote{We stress that while de Sitter group is the (linearly realized) symmetry group of the scalar action, the geometry throughout the galilean genesis phase remains close to flat.} -- the Hubble rate and the sigma model field $e^\pi$ growing as time flows from $t=-\infty$ towards $t=0$. Whenever $e^\pi$ starts exceeding unity on the other hand, the emergent shift symmetry naturally leads to an attractor solution with de Sitter \textit{geometry} on which the scalar acquires a linear profile, $\pi\propto t$ \cite{ArkaniHamed:2003uz, Kobayashi:2010cm}. This qualitatively explains the gradual transformation between genesis and de Sitter phases as illustrated by the blue curve in Fig. \ref{fig:1}. Last but not least, the enhanced symmetries for large and small field values lead to the quantum robustness of the whole qualitative picture. Indeed, both symmetries \eqref{dil} and \eqref{shift} are broken at order one when $e^\pi\sim 1$, making it hard to argue in favor of quantum stability of the detailed intermediate-time behaviour of our solutions. Nevertheless, the scale and shift symmetries are fully intact asymptotically, determining radiative stability of both the early- and the late-time dynamics. Backgrounds exhibiting the genesis-de Sitter transition can thus be expected to exist \textit{generically}, since both of the asymptotic solutions arise solely from symmetry considerations. A similar discussion of quantum robustness has been given in Ref. \cite{Elder:2013gya} in the context of flat-space constructions interpolating between NEC-satisfying and NEC-violating vacua. For completeness, in the rest of the section we give a relatively detailed overview of the two asymptotic regimes of the solutions we wish to study. \subsection*{Galilean genesis} Conformal symmetry, $SO(4,2)$, can be \textit{generically} broken down to its maximal, de Sitter subgroup $SO(4,1)$ by a time-dependent scalar profile \cite{Fubini:1976jm, Nicolis:2008in,Nicolis:2009qm}. One way to achieve such breaking is via the (simplest non-trivial) \textit{conformal galileon} lagrangian \begin{eqnarray} \label{gg} S_{\text{1}}=\int d^4 x \sqrt{-g}~\bigg[ f^2 e^{2\pi}(\partial\pi)^2+\frac{f^3}{\Lambda^3}(\partial\pi)^2\Box\pi+\frac{f^3}{2\Lambda^3} (\partial\pi)^4\bigg ]~. \end{eqnarray} It can be straightforwardly checked that the theory possesses an exact rolling solution on flat spacetime \cite{Nicolis:2009qm,Creminelli:2010ba} \begin{eqnarray} \label{ggsol} e^\pi=-\frac{1}{H_0 t}, \qquad H_0^2=\frac{2\Lambda^3}{3f}~, \end{eqnarray} leading precisely to the $SO(4,2) \rightarrow SO(4,1)$ breaking pattern. The dilatation invariance, left unbroken by the background, leads to vanishing of its energy density\footnote{This immediately follows from scale invariance ($\rho\propto \frac{1}{t^4}$) plus the energy conservation ($\dot\rho=0$).}, $\rho=0$, while the pressure $p=-2f^2/(H_0^2t^4)$ is negative -- implying a strongly NEC-violating ($\dot H\gg H^2$) expansion \cite{Creminelli:2010ba}. The universe described by GG starts out in flat spacetime, the Hubble rate growing according to the second Friedmann equation $2M_{\rm Pl}^2 \dot H=-(\rho+p)$, which upon integration yields \begin{eqnarray} \label{ggsol1} H\simeq-\frac{1}{3}\frac{f^2}{M_{\rm Pl}^2}\frac{1}{H^2_0t^3}~. \end{eqnarray} The time $t_0$ at which gravity starts playing non-negligible role ($H\sim \dot\pi$), can be estimated as in \eqref{tzero}. It roughly coincides with the time of EFT breakdown/start of reheating. Scalar perturbations, relevant for the CMB are instead produced at earlier times $t\mathrel{\rlap{\lower3pt\hbox{\hskip0pt$\sim$} t_0$, via minimally coupling an additional, scaling dimension-0 field $\varphi$ to the 'fake de Sitter' metric $g^{\text{dS}}_{_{\mu \nu}}=e^{2\pi}\eta_{_{\mu \nu}}$. This leads to a scale-invariant spectrum for the spectator $\varphi$ (despite the background metric being practically flat), that can be later imprinted on the physical curvature perturbation $\zeta$ through one of the standard mechanisms \cite{Enqvist:2001zp,Lyth:2001nq,Dvali:2003em}. The near-to-flat geometry on the other hand implies a strongly blue-tilted tensor spectrum $P_h(k)\sim k^2$, largely irrelevant for CMB observations \cite{Creminelli:2010ba}. \subsection*{Galileon inflation} An immediate candidate for describing the late-time de Sitter asymptotics of the solutions of interest is a cubic galileon theory with a small quartic self-interaction, defined by the following action \begin{eqnarray} \label{ginf} S_{\text{2}}=\int d^4 x \sqrt{-g} ~\bigg[f^2 (\partial\pi)^2+\gamma_3\frac{f^3}{\Lambda^3}(\partial\pi)^2\Box\pi+\gamma_4\frac{f^3}{2\Lambda^3} (\partial\pi)^4\bigg ]~, \end{eqnarray} where $\gamma_{3,4}$ are constant parameters. The form of the above action is dictated by the early-time genesis asymptotics. Indeed, both of the interactions in \eqref{ginf} are also present in \eqref{gg}, the only difference between the two theories being that the former lacks scale invariance. Moreover, the galileon term will be crucial for the speed of sound in the inflationary regime to be strictly positive. Inflationary solutions in this theory have been studied in Ref. \cite{Kobayashi:2010cm}. Here we will re-derive all of the (qualitative) results of the latter reference using simple EFT considerations. In addition, we will provide arguments in favor of the quantum robustness of these results -- something that, to the best of our knowledge, has not been pointed out before. The Friedmann equation and the equation of motion for $\pi$ take on the following form on spatially flat FRW backgrounds \begin{gather} H^2=\frac{f^2}{3 M_{\rm Pl}^2 H_0^2} ~\( \gamma_4 \dot\pi^4+4\gamma_3 H\dot\pi^3-H_0^2\dot\pi^2 \)~, \\ \( 4 \gamma_4\dot\pi^2 +8\gamma_3 H\dot\pi -2 H_0^2\)\ddot\pi +4 \gamma_4 H\dot\pi^3+4\gamma_3\(3 H^2+\dot H\)\dot\pi^2-6 H_0^2 H\dot\pi=0 ~, \end{gather} making existence of de Sitter vacua ($H=\text{const}$) with a linear $\pi\propto t$ profile explicit -- a direct consequence of shift-invariance \eqref{shift} of the $\pi$-lagrangian. Furthermore, the expansion rate and the scalar profile can be estimated as \begin{eqnarray} \label{sol1} H^2\sim \frac{f^2}{M_{\rm Pl}^2} H_0^2~,\qquad \dot \pi\sim H_0~. \end{eqnarray} The simplest and the most straightforward way of studying the spectrum of scalar perturbations is based on the effective theory of inflation \cite{Creminelli:2006xe,Cheung:2007st}. The formalism is reviewed in great detail in Appendix \ref{appA} and in Sec.\ref{hdim}, so we will content ourselves with a brief treatment here. The two operators in the effective theory that lead to non-trivial dynamics at high energies are the $\delta N^2$ and $\delta N\delta E^i_{~i}$ terms in the notation of Eq.\eqref{s_pi}. The coefficients of these terms, given for a generalized theory of Sec.\ref{ggal} in Eq.\eqref{M's} (the present case corresponds to simply setting $\mathcal{F}_2=1$ in the latter expressions), are of order \begin{eqnarray} \label{sol2} M^4\sim f^2 H_0^2, \qquad \hat M^3_3\sim f^2 H_0~. \end{eqnarray} A particularly useful regime of the system is the one corresponding to the short-distance, \textit{decoupling} limit, that allows to zoom onto the relevant high-energy degrees of freedom present in the theory\footnote{Although a `short distance' limit, the decoupling limit is crucially valid at distances parametrically greater than the inflationary Hubble scale at which the scalar spectrum is evaluated.}. In this limit the dynamics of the scalar is fully captured by the Goldstone mode corresponding to the breaking of time translation-invariance, which we will refer to as $\pi_g$ (we will rely on the reader to not confuse the Goldstone boson with the fundamental galileon field $\pi$). The decoupling limit action for $\pi_g$, assuming $M_{\rm Pl}^2 \dot H\ll f^2 H_0^2$, reads \cite{Creminelli:2006xe,Cheung:2007st} \begin{eqnarray} \label{goldst} S=\int d^4 x \sqrt{-\bar g}\bigg [M_0^4\(\dot \pi_g^2-c_s^2 \frac{(\nabla\pi_g)^2}{a^2}\)-M_0^4 \dot\pi_g\frac{(\nabla\pi_g)^2}{a^2}+\frac{\hat M^3_3}{2}\frac{\nabla^2\pi_g (\nabla\pi_g)^2}{a^4}+\dots\bigg],~~~~~ \end{eqnarray} where $\bar g$ is the unperturbed de Sitter metric, and we have made use of the following notation \begin{eqnarray} \label{cssq} M^4_0=\frac{M^4}{2}-3\hat M^3_3 H\sim f^2 H_0^2, \qquad c_s^2=\frac{3 \hat M^3_3 H}{M_0^4}~. \end{eqnarray} Recalling that the physical curvature perturbation is related to $\pi_g$ by a gauge transformation, $\zeta=-H\pi_g$, one can directly read off the expression for the power spectrum of scalar perturbations from the Goldstone action \eqref{goldst} \begin{eqnarray} \langle\zeta_{\vec k_1} \zeta_{\vec k_2} \rangle=(2\pi)^3 \delta(\vec k_1+\vec k_2) \frac{1}{k_1^3}~\frac{H^4}{M_0^4 c_s^3}~, \end{eqnarray} where all quantities on the right hand side are assumed to be evaluated at horizon crossing $k_1=a H$, as usual. The tensor spectrum on the other hand is given by the universal formula $\Delta^2_\gamma\sim H^2/M_{\rm Pl}^2$. Using Eqs. \eqref{sol1} and \eqref{sol2}, as well as Eq.\eqref{cssq} for the speed of sound, one finds the following expressions for the dimensionless power spectra \begin{eqnarray} \Delta^2_\zeta \sim \frac{f^{1/2} H_0^2}{M_{\rm Pl}^{5/2}},\qquad \Delta^2_\gamma \sim \frac{f^2 H_0^2}{M_{\rm Pl}^4}, \qquad r=\frac{\Delta^2_\gamma}{\Delta^2_\zeta } \sim \(\frac{f}{M_{\rm Pl}}\)^{3/2}~, \end{eqnarray} in agreement with the results of \cite{Kobayashi:2010cm} (see the latter reference for the computation in the full theory, including the precise numerical factors). Moreover, one can see from the above that the tensor-to-scalar ratio can easily be made large enough to be detectable if $f$ is sufficiently close to $M_{\rm Pl}$ -- all within the regime of validity of the underlying effective field theory. One can go further and estimate the amount of non-Gaussianity that the model under consideration is expected to generate. The most relevant cubic interactions of $\pi_g$, giving the leading non-Gaussian effects have been explicitly written out in \eqref{goldst}. The three-point function is of the equilateral shape for both of these \cite{Bartolo:2010bj} (see also \cite{Kobayashi:2011pc}). The amplitude on the other hand can be estimated e.g. for the $\dot \pi_g (\nabla\pi_g)^2$ operator in the standard way \cite{Cheung:2007st} (again, all terms on the r.h.s. should be understood as evaluated at horizon-crossing) \begin{eqnarray} f_{NL}\sim \frac{1}{\zeta} \frac{\mathcal{L}_{\dot \pi_g (\nabla\pi_g)^2}}{\mathcal{L}_{\dot \pi_g^2}}\sim \frac{1}{H \pi}\frac{H\(k/a\)^2\pi}{H^2}=\frac{1}{c_s^2}~. \end{eqnarray} This leads to the amount of non-Gaussianity similar to that in DBI models of inflation \cite{Alishahiha:2004eh}. An analogous estimate shows that the second cubic self-interaction generates a comparable contribution to $f_{NL}$. \begin{figure} \centering \includegraphics[width=.4\textwidth]{diagram} \caption{The diagram, responsible for the dominant quantum correction to the background solution in galileon inflation.} \label{diagram} \end{figure} At this point one may be worried about the UV sensitivity of the obtained inflationary solution, since the scales suppressing the two (canonically-normalized) interactions in \eqref{ginf}, $\Lambda$ and $\tilde \Lambda\equiv (f\Lambda^3)^{1/4}$, are parametrically separated\footnote{The same is true for the conformally invariant theory \eqref{gg}, however there the hierarchy is not a problem, as it is completely stabilized by conformal symmetry.} ($\tilde\Lambda\gg \Lambda$). In fact, this separation is crucial if all three terms in the lagrangian are to play an equally important role on the given background; indeed, for $f\simM_{\rm Pl}$, one can estimate the magnitude of each operator (including the kinetic term) to be of order $\rho_{dS}\equiv f^2 H_0^2$. Interpreting $\Lambda$ -- the smallest of the two scales -- as the quantum cutoff of the theory then, nothing apparently prevents a loop-generated self-interaction \textit{e.g.} of the form $(\partial\pi)^4/\Lambda^4$, which would parametrically dominate over the last term (and therefore over all terms) in \eqref{ginf}. This would impair the whole description of the obtained dS backgrounds. Fortunately, the latter reasoning turns out to be too hasty and the background can in fact be trusted. This can be seen as follows. Consider all loop diagrams, generating a term of the form $(\partial\pi)^{2n}$. What is the smallest scale that can suppress such an operator? To answer this question, we note that whatever the diagram responsible for this operator is, it can not have an external leg originating from the cubic galileon vertex, since this would lead to at least two derivatives acting on the corresponding asymptotic state (the reason for this lies in the non-renormalization theorem that severely constrains the form of quantum corrections in galileon theories \cite{Luty:2003vm}). We thus conclude that all external legs in the diagram originate from the quartic interaction, which introduces a suppression of \textit{at least one} factor of $f$ per pair of fields in the corresponding effective vertex. The least suppressed loop corrections of the given form thus correspond to the diagram of Fig. \ref{diagram}. Assuming that all the rest of the vertices are those of the cubic galileon (and therefore only introduce factors of $\Lambda$, but not of $f$) and that loop integrals are cut-off at energies of order the strong-coupling scale of the theory $\Lambda$, one arrives at the following conservative estimate for the magnitude of the operators of the given type \begin{eqnarray} \mathcal{L}_{loop}= \frac{(\partial\pi)^{2n}}{f^n \Lambda^{3k}}, \qquad k=n-\frac{4}{3}~. \end{eqnarray} Evaluating $\mathcal{L}_{loop}$ on the classical de Sitter background gives \begin{eqnarray} \mathcal{L}_{loop}=f^{4/3} H_0^{8/3}\ll \rho_{ds}~, \end{eqnarray} independently of $n$. This leads one to conclude that quantum corrections of the form $(\partial\pi)^{2n}$ do not modify the background obtained from the lagrangian (\ref{ginf}). Note that the non-renormalization properties of the galileon play a crucial role in the latter conclusion. Furthermore, the fact that $\pi$ acquires a linear profile on de Sitter backgrounds makes operators with more than one derivative per field similarly irrelevant, since they are suppressed by powers of the scale $H_0$, parametrically smaller than $f$ and $\Lambda$. The $\pi\propto t$ solution describes a perfect de Sitter space, leading to exactly scale-invariant perturbations; adding a small potential (or deforming the form of the action otherwise), both the scalar and the tensor modes can be produced with slightly tilted spectra -- just as they are in the canonical inflationary case. In addition, to complete the picture one of course has to specify a mechanism for exiting the de Sitter phase. There are known ways of achieving this, and we refer the interested reader to works, dealing with similar issues in various contexts \cite{ArkaniHamed:2003uz, Senatore:2004rj, Osipov:2010ee, Ivanov:2014yla}. \subsection*{An explicit example} As a simple example of a theory with the above-described asymptotics, one can consider the deformed galilean genesis lagrangian \begin{eqnarray} S=\int d^4x~ \sqrt{-g} ~\bigg[\frac{1}{2}M_{\rm Pl}^2 R +f^2~ \frac{e^{2\pi}}{1+\beta e^{2\pi}} ~(\partial\pi)^2+\frac{f^3}{\Lambda^3}(\partial\pi)^2\Box\pi +\frac{f^3}{2\Lambda^3}(\partial\pi)^4 \bigg]~, \label{simplemodel} \end{eqnarray} with $\beta$ an arbitrary constant. For $\beta=0$ the theory is just the conformal galileon and when starting out in the GG phase, the expansion rate of the universe diverges and the background exits the regime of validity of the EFT at some finite time (see the red curve in Fig. \ref{fig:1}) -- the scalar profile growing as $e^\pi\sim 1/t$ throughout. For a nonzero $\beta$ on the other hand, the dynamics of the system is completely altered as soon as $\beta e^{2\pi}$ becomes of order, or greater than one: the theory becomes effectively described by a $P(X)$ - type lagrangian with a cubic galileon self-interaction, resulting in transition into an inflationary de Sitter phase. The corresponding solutions are studied in Appendix \ref{num}, where the existence of extended genesis cosmologies is illustrated via numerical analysis: the system clearly exhibits transition from genesis into a quasi - de Sitter regime precisely around the time $t_0$ given in \eqref{tzero}, see Fig. \ref{figuretwo}. Perhaps the only downside of this simple model is the short temporal region with gradient instability at intermediate times: while completely free from ghosts, the squared speed of sound of the scalar perturbation goes slightly negative on the given background around $t\sim t_0$ for a period of roughly a Hubble time, as shown in Fig. \ref{figuretwo} (we will track down the origin of the gradient instability analytically in Sec. \ref{analytic} ). While certainly a problem in the classical theory, higher-order effects can in principle take care of this issue -- rendering the cosmological evolution free from instabilities, see the discussion in Sec. \ref{hdim} and Appendix \ref{num}. We refer the reader to Appendix \ref{num} for a detailed discussion of numerical solutions to the illustrative model \eqref{simplemodel}, and turn to a systematic construction of theories leading to early universe cosmology with the genesis - dS transition in the next section. \section{Generalized galileons} \label{ggal} In the present and the next sections we will take on the task of obtaining (analytic) cosmological solutions exhibiting extended genesis. Rather than constructing solutions to a particular theory obeying the asymptotic scale and shift symmetries described in Sec. \ref{sec2}, we will employ the trick used in Ref. \cite{Elder:2013gya}, where the appropriate theory itself is inverse - engineered based on a postulated ansatz for the desired cosmological solution. The asymptotic symmetries, as we will see, then follow automatically from the construction which we describe in what follows. Consider a (generally dilatation-breaking) deformation of the galilean genesis lagrangian \begin{eqnarray} \label{ggg} \mathcal{S}_\pi=\int d^4 x ~\sqrt{-g} ~\bigg[f^2 \mathcal{F}_1(\pi) (\partial\pi)^2+\frac{f^3}{\Lambda^3} (\partial\pi)^2\Box\pi +\frac{f^3}{2\Lambda^3} \mathcal{F}_2(\pi) (\partial\pi)^4 \bigg ] \end{eqnarray} where $\mathcal{F}_{1,2}$ are \textit{a priori} arbitrary dimensionless functions of the galileon field $\pi$. We will interchangeably use the two scales $\Lambda$ and $H_0$ (as defined in \eqref{ggsol}) throughout. The dynamics of the system is governed by the Einstein's equations plus the scalar equation of motion. These however are not independent: as a consequence of diffeomorphism invariance, the scalar equation can be traded for the conservation of its stress-energy tensor via \begin{eqnarray} \label{s.eom} \nabla_\mu T^\mu_{~\nu}=-\frac{\delta S}{\delta \pi} \partial_\nu\pi~. \end{eqnarray} On homogeneous FRW backgrounds, it is the energy conservation, $\dot \rho + 3H (\rho+p)=0$, that yields the $\pi$ equation of motion. Energy conservation on the other hand follows from the temporal and space components of the Einstein's equations -- therefore we can choose the latter two to make up a complete system determining background evolution. The stress-energy tensor, sourced by $\pi$ in \eqref{ggg} is \bea T^\pi_{_{\mu \nu}}=&-&f^2 \mathcal{F}_1(\pi) ~[2\partial_\mu\pi\partial_\nu\pi-g_{_{\mu \nu}} (\partial\pi)^2]\nonumber \\ &-&\frac{f^3}{\Lambda^3}~[2\partial_\mu\pi\partial_\nu\pi\Box\pi- \partial_\mu\pi\partial_\nu(\partial\pi)^2-\partial_\nu\pi\partial_\mu(\partial\pi)^2 +g_{_{\mu \nu}}\partial_\lambda\pi\partial^\lambda(\partial\pi)^2]\nonumber \\ &-&\frac{f^3}{2 \Lambda^3}~\mathcal{F}_2(\pi) ~[4 (\partial\pi)^2\partial_\mu\pi\partial_\nu\pi-g_{_{\mu \nu}} (\partial\pi)^4]~, \end{eqnarray} leading to the following expressions for the energy density and pressure due to a homogeneous $\pi$-profile \bea \label{rho} \rho &=&\frac{f^2}{H_0^2} ~\dot \pi^2 \big [ \mathcal{F}_2(\pi)\dot\pi^2+4 H\dot\pi - H_0^2\mathcal{F}_1(\pi)\big ] ~,\\ \label{press} p &=& \frac{f^2}{3 H_0^2}~\dot\pi^2 \bigg[ \mathcal{F}_2(\pi) \dot\pi^2-4\ddot \pi-3H_0^2\mathcal{F}_1(\pi) \bigg]~. \end{eqnarray} The two functions $\mathcal{F}_{1,2}(\pi)$ can be solved for with the help of the temporal and spatial components of Einstein's equations, $3 M_{\rm Pl}^2 H^2=\rho$ and $M_{\rm Pl}^2 (3 H^2+2 \dot H)=-p$, which yields \begin{eqnarray} \label{f1} \mathcal{F}_1&=&\frac{6 M_{\rm Pl}^2 H_0^2 H^2+3M_{\rm Pl}^2 H_0^2\dot H-2 f^2 H\dot \pi^3-2f^2 \dot\pi^2\ddot\pi}{f^2 H_0^2\dot\pi^2}\\ \label{f2} \mathcal{F}_2&=&\frac{9M_{\rm Pl}^2 H_0^2 H^2+3M_{\rm Pl}^2 H_0^2\dot H-6 f^2 H\dot \pi^3-2f^2 \dot\pi^2\ddot\pi}{f^2 \dot\pi^4}~. \end{eqnarray} Now, for any \textit{postulated} homogeneous profile of the scalar and the Hubble rate, one can find the theory (i.e. find $\mathcal{F}_{1,2}(\pi)$) such that the desired background solves its equations of motion. The recipe for constructing the relevant solutions is given as follows: \begin{itemize} \item Postulate background profiles $\pi_0(t)$ and $H(t)$~ \item For the chosen background solutions, find the time-dependent functions $\mathcal{F}_{1,2} (t)$ with the help of \eqref{f1} and \eqref{f2} \item Invert the expression for $\pi_0(t)$ to find $t=t(\pi_0)$ \item Using the previous steps, find $\mathcal{F}_{1,2}$ as functions of $\pi_0$: $\mathcal{F}_{1,2}=\mathcal{F}_{1,2}\(t(\pi_0)\)$~. \end{itemize} That way one can formally construct theories admitting arbitrary cosmological profiles for $\pi$ and $H$. Although such an \textit{ad hoc} construction might look uncomfortable, we will see that at least for the solutions we will be interested in, it will lead to theories that enjoy various types of asymptotic symmetry, making them highly non-generic in the sense discussed in Sec. \ref{sec2}. \subsection*{Perturbations} As a next step, we check whether the cosmological solutions obtained through the above procedure are stable. This can be done with the help of the analysis spelled out in Appendix \ref{appA}. In the unitary gauge, defined by the absence of $\pi$ - fluctuations, $\pi(x,t)=\pi_0(t)$, the only scalar degree of freedom present in the theory is captured by the standard curvature perturbation of equal-density hypersurfaces $\zeta$, that enters into the perturbed spatial metric in the following way \begin{eqnarray} \label{gij} g_{ij}=a(t)^2 (1+2\zeta)\delta_{ij}~. \end{eqnarray} The curvature perturbation is an exactly massless field, which directly follows from the fact that $\zeta =const$ should be a legitimate solution, since $g_{ij}$ in this case is obtained from the unperturbed FRW metric by a mere constant rescaling of spatial coordinates (this, of course, is also the origin of conservation of $\zeta$ at super-horizon distances). Having the background quantities at hand, one can readily derive the quadratic $\zeta$ action following the standard procedure \cite{Maldacena:2002vr} \begin{eqnarray} \label{quadact} S_\zeta=\int d^4x~ a^3~\bigg[A(t)~\dot\zeta^2-B(t)~\frac{1}{a^2}\(\vec{\nabla}\zeta\)^2-C(t)~\frac{1}{a^4}\(\vec{\nabla}^2\zeta\)^2 \bigg]~. \end{eqnarray} The kinetic coefficients $A$ and $B$ are found to be \cite{Creminelli:2006xe,Creminelli:2010ba} \bea \label{A} A(t) &=&\frac{M_{\rm Pl}^2 (-4 M_{\rm Pl}^4 \dot H-12M_{\rm Pl}^2 H \hat M^3+3\hat M^6+2M_{\rm Pl}^2 M^4)}{(2M_{\rm Pl}^2H-\hat M^3)^2}~,\\ \label{B} B(t)&=&\frac{M_{\rm Pl}^2 \(-4 M_{\rm Pl}^4 \dot H+2M_{\rm Pl}^2 H \hat M^3-\hat M^6+2M_{\rm Pl}^2\partial_t\hat M^3\)}{(2M_{\rm Pl}^2H-\hat M^3)^2}~, \end{eqnarray} while $C(t)=0$ for our `classical' action \eqref{ggg} (it will be nonzero once we include higher-order terms in the effective theory in Sec. \ref{hdim}). Explicit expressions for the time-dependent coefficients $\hat M^3$ and $M^4$ are given in Eq. \eqref{M's}. Apart from other background quantities, these explicitly depend on the function $\mathcal{F}_2(\pi_0)$. Using the expression \eqref{f2} for the latter, one finds \begin{eqnarray} \label{Anec} A&=&3 M_{\rm Pl}^2 ~\frac {36 M_{\rm Pl}^4 H_0^4 H^2+9M_{\rm Pl}^4 H_0^4\dot H-18 M_{\rm Pl}^2 f^2 H_0^2H \dot\pi ^3 -6M_{\rm Pl}^2 f^2 H_0^2 \dot\pi ^2\ddot\pi+4 f^4 \dot\pi^6}{(3M_{\rm Pl}^2 H_0^2 H-2f^2 \dot\pi^3)^2},~~~~~~~~\\ \label{Bnec} B&=&\frac{-9M_{\rm Pl}^6 H_0^4 \dot H+6M_{\rm Pl}^4 f^2H_0^2H\dot\pi^3+18M_{\rm Pl}^4 f^2 H_0^2 \dot \pi^2\ddot\pi -4M_{\rm Pl}^2 f^4\dot\pi^6}{(3M_{\rm Pl}^2 H_0^2 H-2f^2 \dot\pi^3)^2}, \end{eqnarray} while the speed of sound for short wavelength scalar perturbations is given by $c_s^2=A/B$. Positive $A$ and $B$ throughout the entire course of cosmological evolution guarantee the absence of ghost and gradient instabilities respectively. As a quick check, one can apply the above piece of formalism to galilean genesis \cite{Creminelli:2010ba}. Plugging the scalar and Hubble profiles, \eqref{ggsol} and \eqref{ggsol1} into the expressions for the curvature perturbation's kinetic coefficients \eqref{Anec} and \eqref{Bnec}, one obtains the following values for the latter quantities to the leading order in $M_{\rm Pl}$: \begin{eqnarray} A(t)=B(t)=\frac{9M_{\rm Pl}^4H_0^2}{f^2} ~t^2~. \end{eqnarray} This precisely agrees with the expressions found in \cite{Creminelli:2010ba}. \section{Extended genesis: analytic solutions} \label{analytic} While the recipe, spelled out in the previous section formally allows to construct theories admitting essentially arbitrary cosmological solutions, most of these fail to be physically meaningful in one way or another. A generic such solution will lead to either ghost or gradient instability at the level of small perturbations; moreover, most of the resulting theories will be free from symmetries -- even the asymptotic ones, casting shadow on quantum robustness of the whole picture. Nevertheless, we will show in this section that a class of theories exists, that admit completely stable cosmological solutions interpolating between a low/zero curvature maximally symmetric spacetime in the far past and a larger curvature inflationary dS spacetime in the future -- with a strong/moderate violation of the null energy condition in between. Importantly, we will see that asymptotically these theories enjoy symmetries of the kind described in Sec. \ref{sec2}. Let us work in a coordinate system such that time runs from $t =-\infty$ towards $t=0$ over the cosmological phase of interest. At (or shortly after) $t=0$, the system is assumed to reheat, or exit the given phase otherwise. Inspired by the early-time galilean genesis asymptotics \eqref{ggsol} and \eqref{ggsol1}, we will adopt the following ansatz for the Hubble rate \begin{eqnarray} \label{ansatz} H=\lambda +\beta ~\frac{f^2}{M_{\rm Pl}^2 H_0^2} ~\dot\pi_0^3~, \end{eqnarray} where $\lambda$ and $\beta$ are free parameters (of mass dimension one and zero respectively) of the theory, giving rise to the solution of interest. For the scalar, we will assume the ansatz of the following form (which is again motivated by the genesis solution) \begin{eqnarray} \label{ansatz1} e^{\pi_0}=\frac{1}{H_0} \frac{1}{t_*-t}~. \end{eqnarray} Here, $t_*>0$ is yet another free parameter with mass dimension minus one. While resembling GG at early times (and for sufficiently small $\lambda$), \eqref{ansatz} and \eqref{ansatz1} describe a cosmology regularized towards $t\to 0^-$, so that none of the invariants in the theory grow unbounded over the entire interval $t\in [-\infty,~0]$. Galilean genesis is recovered at all times for the particular values of the parameters $\lambda=0$, $\beta=1/3$ and $t_*=0$. For $\lambda\neq 0$ on the other hand, there is a crucial difference: rather than from flat, Minkowski spacetime, the system starts out evolving from de Sitter space with the curvature set by the parameter $\lambda$. In order for the universe to be described by inflationary de Sitter geometry at $t\to 0^-$, the parameters of the theory should satisfy certain constraints. One such constraint arises from requiring the Hubble rate not to vary considerably over a single e-fold at $|t|\ll t_*$. The necessary condition for that is: \begin{eqnarray} \label{dscond} 1\gg \varepsilon \equiv \frac{\dot H}{H^2}\bigg |_{t\to 0}\sim \begin{cases} \frac{M_{\rm Pl}^2 H_0^2}{\beta f^2} ~t_*^2~, &\text{if} ~~\lambda \ll \beta~\frac{ f^2}{M_{\rm Pl}^2 H_0^2}~\frac{1}{ t_*^3} \\ \frac{\beta}{\lambda^2}~ \frac{f^2}{M_{\rm Pl}^2 H_0^2}~\frac{1}{ t_*^4}~, &\text{if} ~~ \lambda \gg \beta~\frac{ f^2}{M_{\rm Pl}^2 H_0^2}~\frac{1}{ t_*^3}~. \end{cases} \end{eqnarray} Not surprisingly, this condition is equivalent to the one constraining $\dot \pi$ to be quasi-constant at late times: \begin{eqnarray} \frac{1}{H}\frac{d}{dt} \ln \dot\pi_0\ll 1 ~. \end{eqnarray} This shows that $\pi$ can indeed be approximated by a linear profile towards $t\to 0^-$, leading to galileon inflation discussed in Sec. \ref{sec2}. In the rest of this section we will study various interesting regions in the six-dimensional space spanned by the free parameters $\( M_{\rm Pl}, f,H_0,\lambda,\beta, t_* \)$ of the theory. \subsection{$\lambda =0$} We begin with the case that, in the asymptotic past, the system starts out evolving from flat spacetime. This happens for $\lambda=0$. As a quick consistency check, one can \textit{derive} the conformally invariant GG lagrangian \eqref{gg} from our ansatz for the extended genesis cosmology, following the inverse construction of the previous section. Indeed, plugging \eqref{ansatz} and \eqref{ansatz1} (with $\beta=1/3$) into the expressions for $\mathcal{F}$-functions, \eqref{f1} and \eqref{f2}, we find at the leading order in $1/M_{\rm Pl}^2$ (and at times $|t|\gg t_*$) \begin{eqnarray} \mathcal{F}_1=\frac{1}{H_0^2 t^2}=e^{2\pi}, \qquad \mathcal{F}_2=1~. \end{eqnarray} This precisely corresponds to the conformal galileon. For values of $\beta$ other than $1/3$, on the other hand, our ansatz describes subluminal versions of GG \cite{Creminelli:2012my} at $|t|\gg t_*$. Concentrating on the full solution, including times $|t|\leq t_*$, stability of the system requires that the kinetic coefficients in \eqref{quadact} are positive at all times. For $\lambda=0$, they are given as follows \begin{eqnarray} \label{atilde} \frac{3 (2M_{\rm Pl}^2 H-\hat M^3)^2 H_0^4}{4M_{\rm Pl}^2}~A&=&2(4+15 x +18x^2) f^4 \dot\pi^6+3(4+9x)M_{\rm Pl}^2 f^2 H_0^2\dot\pi^2\ddot\pi~,~~~~~\\ \label{btilde} \frac{3 (2M_{\rm Pl}^2 H-\hat M^3)^2 H_0^4}{4M_{\rm Pl}^2}~B&=& \(2 x f^2\dot\pi^4-9x M_{\rm Pl}^2 H_0^2\ddot\pi\)f^2 \dot\pi^2~, \end{eqnarray} where we have defined $x=\beta-2/3$ for further convenience. As an immediate observation, we note that $A$ is manifestly positive for positive $x$ (both $\dot\pi$ and $\ddot\pi$ are positive at all times for our ansatz), while $B$ does not have a definite sign. For the special case that the parameter $x$ is small however, $B$ can be made arbitrarily small, compared to $A$, implying a vanishing speed of sound for $\zeta$. This is similar to what happens in ghost condensation, where the absence of gradient instability is determined by higher-order operators in the effective theory. It is straightforward to see that $B$ cannot be positive over the entire temporal interval of interest -- at least for our ansatz \eqref{ansatz}. Indeed, we are interested in solutions, that start in galilean genesis at $t\to -\infty$ and end up in the inflationary phase at $t\to 0^-$. As shown in the previous section, the latter phase requires $\dot\pi$ to be practically constant, meaning that the second term in the parentheses on the r.h.s. of \eqref{btilde} should be negligible compared to the first one at late times. Positivity of $B$ at late times then requires $x>0$. On the other hand, galilean genesis corresponds to the second term prevailing at sufficiently early times, since $\ddot \pi\sim 1/t^2$ decreases parametrically slower than $\dot\pi^4\sim 1/t^4$ at large and negative $t$. For $x>0$ however, this would lead to gradient instability at early times. In contrast, in the opposite case of $x<0$, one would recover gradient instability at late times, while the early-time genesis phase would be completely stable. One is therefore led to conclude that gradient instability is unavoidable for the given choice of the ansatz \eqref{ansatz} in the $\lambda=0$ case -- at least at the leading order in derivative expansion. Concentrating on negative $x$ (so that the genesis phase is stable), the time at which gradient instability occurs (i.e. when $B$ flips sign) is of order $|\tau|\sim f/(M_{\rm Pl} H_0)$. The slow-roll parameter at that time can be readily estimated, $\varepsilon\sim M_{\rm Pl}^2 H_0^2 \tau^2/f^2\sim 1$, see Eq. \eqref{dscond}. This means that the gradient instability for $\lambda=0$ solutions necessarily kicks in before the onset of the de Sitter regime, explaining the pattern we have found via numerical analysis in Sec. \ref{sec2} (see also Appendix \ref{num}). We end the present subsection with a couple of consistency checks for our calculations. First, we note that for $x=-1/3$ corresponding to galilean genesis, one recovers an exactly luminal scalar mode, $c_s^2=\tilde B/\tilde A=1$ at early times. Moreover, as stressed several times above, the late-time de Sitter phase should correspond to an enhanced shift symmetry on $\pi$. That this is indeed the case is the result of quasi-constancy of the $\mathcal{F}$ functions \begin{eqnarray} \frac{1}{H} ~\frac{d}{dt}\ln \mathcal{F}_{1,2} \ll 1~. \end{eqnarray} which, as can be straightforwardly verified, directly follows from (the $\lambda=0$ version of) Eq. \eqref{dscond} -- the condition for the universe to be described by de Sitter geometry at $|t|\ll t_*$. \subsection{$\lambda \neq 0$} We now turn to the case that in the asymptotic past the universe starts out evolving from de Sitter space, rather than Minkowski, $\lambda\neq 0$. The curvature of the initial state is of order $R\sim \lambda^2$ and is a free parameter of the theory; if its value is strictly zero, we have seen that the resultant cosmological solution suffers from a gradient instability before the onset of de Sitter regime for much of the parameter space -- at least if one ignores higher-order operators in the effective theory. However, for non-zero $\lambda$, as we will now demonstrate, gradient instabilities can be avoided even in the 'classical' theory, that is without invoking higher-derivative terms in the EFT for perturbations. The kinetic coefficients \eqref{Anec} and \eqref{Bnec}, evaluated on the given ansatz are: \begin{eqnarray} A&=&\frac{M_{\rm Pl}^2}{3}~ \frac{36 M_{\rm Pl}^4 H_0^4 \lambda^2\tau^6+3 M_{\rm Pl}^2 f^2 H_0^2 [(-(10+24 x)\lambda\tau +4+9x )]\tau ^2+f^4 (8+30x+36 x^2)}{(f^2 x-M_{\rm Pl}^2 H_0^2\lambda \tau^3)^2}~,\nonumber \\ B&=&\frac{M_{\rm Pl}^2}{3}~ \frac{M_{\rm Pl}^2 f^2 H_0^2 (-2\lambda\tau-9x)\tau^2 + 2 f^4 x}{(f^2 x-M_{\rm Pl}^2 H_0^2\lambda \tau^3)^2}~\nonumber , \end{eqnarray} where we have defined $\tau\equiv t-t_* \leq -t_*$ . An important observation that we will use in what follows is that for positive $x$, and for $\bar \varepsilon \equiv \lambda t_*>9 x/2$, both $A$ and $B$ are manifestly positive (and finite) \textit{at all times}, as can be readily verified by inspecting the above expressions. Given that a strictly vanishing $\lambda$ is not allowed by stability, how small can it be? The smallness of the initial curvature can be conveniently characterized by \begin{eqnarray} \label{hsep} \frac{H(t=0)}{H(t=-\infty)}=\(\lambda~\frac{M_{\rm Pl}^2 H_0^2 t_*^3}{\beta f^2}\)^{-1}\sim \frac{1}{\varepsilon \bar\varepsilon}~. \end{eqnarray} Note that, while $\varepsilon\ll 1$ is required by the late-time de Sitter space, $\bar \varepsilon$ is in principle an unconstrained parameter of the theory. To summarize, choosing $x< 2\bar \varepsilon/9$, one can arrange for a manifestly stable cosmological solution, interpolating between two de Sitter spacetimes with an arbitrary ratio of the corresponding asymptotic curvatures. Moreover, the larger is the separation between the asymptotic Hubble rates \eqref{hsep}, the smaller is the deviation of the late-time geometry from perfect de Sitter space. The speed of sound of the curvature perturbation at $t=0$ can be readily evaluated from the above expressions for the kinetic coefficients \begin{eqnarray} \label{cssq1} c_s^2(t=0)=\frac{x (2-9 \varepsilon)+2\varepsilon \bar\varepsilon}{8+30 x+36 x^2+\mathcal{O}(\varepsilon)}~. \end{eqnarray} Note that the asymptotic $c_s^2$ is finite. For $\varepsilon=0$, its magnitude is bounded from above by $c_s^2<0.031$, which can be found by maximizing the expression \eqref{cssq1} for the squared speed of sound\footnote{Cf. the analytic bound on the scalar speed of sound $c_s^2<0.031$ in galileon inflation, quotted in \cite{Kobayashi:2010cm}.}. Let us for simplicity set $x= 0$ from now on. One distinct property of our ansatz is that the coefficient $A$, having a contribution constant in time, becomes parametrically greater than $B$ at $|t|\gg t_0$, as $B\sim -1/t^3$ at large and negative $t$. This means that the speed of sound of the curvature perturbation tends to zero at early times. What is the theory describing the asymptotic past of the background solutions at hand? To answer this question, we evaluate the $\mathcal{F}$ functions from our deformed galileon action \eqref{ggg}. At the leading order in $1/t$, one finds \begin{eqnarray} \mathcal{F}_1&=&6~\frac{M_{\rm Pl}^2 \lambda^2}{f^2 H_0^2}~ (H_0 t)^2=6~\frac{M_{\rm Pl}^2 \lambda^2}{f^2 H_0^2}~e^{-2\pi}~,\nonumber \\ \mathcal{F}_2&=&9~\frac{M_{\rm Pl}^2 \lambda^2}{f^2 H_0^2}~ (H_0 t)^4=9~\frac{M_{\rm Pl}^2 \lambda^2}{f^2 H_0^2}~e^{-4\pi}~,\nonumber \end{eqnarray} which implies the following form of the scalar action \begin{eqnarray} \label{earlydslag} \mathcal{S}^{\text{early}}_\pi=\int d^4 x ~\sqrt{-g} ~\bigg[ 6~\frac{M_{\rm Pl}^2 \lambda^2}{ H_0^2}~e^{-2\pi} (\partial\pi)^2+\frac{2}{3}~\frac{f^2}{H_0^2} (\partial\pi)^2\Box\pi +3~\frac{M_{\rm Pl}^2 \lambda^2}{H_0^4}~e^{-4\pi} (\partial\pi)^4 \bigg ]~.\nonumber \end{eqnarray} In the regime of interest, $e^\pi\sim1/t$ and the first and the third terms in the parentheses are constant, while the second (the cubic galileon) goes as $\sim 1/t^3$ and is thus completely irrelevant in the asymptotic past\footnote{The latter estimate comes from the $H \dot \pi^3 $ piece, coming from the expansion of the covariant derivative on a de Sitter background.}. \begin{figure} \centering \includegraphics[width=.46\textwidth]{fig3_eps} \quad \includegraphics[width=.46\textwidth]{fig3_cssq} \caption{The 'slow-roll' parameter $\varepsilon$ (left) and the speed of sound of the curvature perturbation $c_s^2$ (right) as functions of time on the solution \eqref{ansatz}, \eqref{ansatz1}. The scales $f$ and $M_{\rm Pl}$ have been assumed equal, while the rest of the parameters have been chosen to be: $H_0=1, ~\lambda=10^{-3},~ t_*=10^{-2},~ x=0$. The two colors correspond to $\varepsilon <10$ (blue) and $\varepsilon>10$ (red).} \label{figurethree} \end{figure} Once the cubic galileon is neglected however, the theory acquires a global symmetry. To see it, it is useful to define a new field $\chi=e^{-\pi}$, in terms of which the two relevant operators are simply $(\partial\chi)^2$ and $(\partial\chi)^4$, and the new symmetry is immediately identified as invariance under constant shifts $\chi\to \chi+c$ (while in terms of $\pi$ this symmetry looks more complicated: $\pi\to -\ln \(e^{-\pi}+c\)$). This shows, that the early-time $\lambda\neq 0$ cosmology is effectively described by a ghost condensate - type theory, albeit written in obscure variables (and hence the vanishing speed of sound)! Needless to say, the emergent global symmetry comes hand-in-hand with all the attractive properties, classical or quantum, characteristic of ghost condensation\footnote{That the given solution indeed describes ghost condensation can also be seen from the fact that $\chi$ acquires a linear profile, $\chi=-H_0 t$, just as the ghost field does on self-accelerated backgrounds.}, see \cite{ArkaniHamed:2003uy,ArkaniHamed:2003uz}. To get a more quantitative perspective on the above discussion, let us consider the solutions \eqref{ansatz} and \eqref{ansatz1} for a specific set of available parameters. As an immediate observation, we note that the Hubble rate does not depend on the magnitude of $f$ and $M_{\rm Pl}$ separately (as far as external matter or bare cosmological constant are not introduced into the system) -- physical quantities are only sensitive to the ratio of the two scales. As a result, one can arbitrarily set the physical units for any one quantity at any one instant of time. For example, the Hubble scale at time $t=0$ can be freely chosen to be $H(0)=10^{14} ~GeV$ in some putative system of units where $H_0\equiv 1$. With this in mind, we set $f=M_{\rm Pl}$, and consider the following values for the rest of the parameters: $H_0=1, ~\lambda=10^{-3},~ t_*=10^{-2},~ x=0$, satisfying (the first case of) the late-time dS condition, Eq. \eqref{dscond}. The time-dependence of the `slow roll' parameter $\varepsilon=\dot H/H^2$ (left) and the speed of sound of the curvature perturbation $c_s^2$ (right) for the above choice of the theory parameters is shown in Fig. \ref{figurethree}. From how $\varepsilon$ depends on time, one can distinguish three stages of evolution, according to whether the system violates the NEC strongly (red), or weakly (blue). The universe starts out in de Sitter space ($\varepsilon \simeq 0$) with tiny curvature $\sim \lambda^2$, the Hubble rate as well as the slow-roll parameter $\varepsilon$ gradually increasing with time. When $\varepsilon \simeq 10$, it enters into the galilean genesis phase with strong violation of the null energy condition. Peaking at $\varepsilon \sim 10^2$ at intermediate times, NEC-violation weakens down back to $\varepsilon \simeq 10$ at $t\simeq -1.5$ (signalling the beginning of the third, galileon inflation stage), $\varepsilon$ decreasing to sub per-cent values shortly afterwards (the final phase of the system corresponds to the blue ends of the curves near $t\to 0$ in Fig. \ref{figurethree}). While the concluding, inflationary de Sitter phase seems rather short in its extension in time, the large magnitude (in units of $H_0$) of the expansion rate at those times allows it to accomodate a large number of e-folds. Indeed, from $t=-0.1$ ($\varepsilon \simeq 5\cdot 10^{-2}$) up until $t=0$ ($\varepsilon \simeq 5\cdot 10^{-4}$), the number of times the scale factor doubles can be easily estimated \begin{eqnarray} N_e=\int\limits_{-0.1}\limits^{0} H dt\simeq 3300~, \end{eqnarray} showing that the de Sitter phase towards the end of the temporal interval of interest is in fact very extended. Furthermore, the Hubble parameter at $t=0$ is $H(0)\sim 10^6$, implying a huge ratio of de Sitter expansion rates in the asymptotic future and the asymptotic past \begin{eqnarray} \frac{H(0)}{H(-\infty)}\sim 10^9~. \end{eqnarray} The right panel of Fig. \ref{figurethree} shows the evolution of the scalar speed of sound. As remarked above, $c_s^2$ starts evolving from nearly zero value at early times, as required by ghost condensate-type cosmologies. Peaking at $c_s^2\simeq 2\cdot 10^{-3}$ during the genesis stage, it drops down again towards late-time galileon inflation. While ghost condensation, described by a $P(X)$-type theory implies vanishing speed of sound of the scalar perturbation at the leading order \cite{ArkaniHamed:2003uy}, galileon inflation (described by a $P(X)$ lagrangian plus one or more galileon terms) does not necessarily lead to $c_s^2=0$ although, as discussed before, there is an upper bound $c_s^2\leq 0.031$ in the latter class of models with a single cubic galileon \cite{Kobayashi:2010cm}. Our solutions however qualitatively (and crucially) differ from 'tilted' ghost condensate with NEC violation considered in \cite{Creminelli:2006xe} in that the speed of sound, although small, is strictly \textit{positive} at all times. The latter is not true for pure $P(X)$ theories: violation of the null energy condition unambiguously implies gradient instabilities at the leading order in the ghost condensate \cite{Hsu:2004vr,Dubovsky:2005xd}. At early times, the tiny speed of sound of the scalar mode suggests that higher-order operators in the effective theory for perturbations \cite{Creminelli:2006xe,Cheung:2007st} could be qualitatively affecting the dynamics of the system. Moreover, depending on the nature of the UV completion, higher-derivative terms could also play a role in the intermediate, galilean genesis phase. In order to estimate these effects, we turn to exploring the structure of the next-to-leading-order action in the EFT formalism in the following section. \section{Beyond the leading order} \label{hdim} The tiny asymptotic scalar speed of sound found for the EG solutions motivates to go beyond the leading order in the EFT for perturbations to assess the role of higher-derivative operators in stability of the system. The generic action for metric fluctuations on a FRW background driven by a single `clock' has the following form (excluding the Einstein-Hilbert part) \cite{Creminelli:2006xe,Cheung:2007st} \begin{eqnarray} \label{s_pi} S_\pi&=&\int d^4x~\sqrt{g_3}N\bigg[-M_{\rm Pl}^2 \dot H \frac{1}{N^2}-M_{\rm Pl}^2(3 H^2+\dot H)\nonumber \\ &+&\frac{1}{2} M^4(t) (\delta N)^2-\hat M_3^3(t)\delta E^i_{~i}\delta N -\frac{\bar M'(t)^2}{2}\delta E^{ij} \delta E_{ij}-\frac{\bar M(t)^2}{2}\delta E^{i~2}_{~i}+\dots \bigg], \end{eqnarray} where $g_3,~N$ and $N_i$ are the standard ADM variables \cite{Arnowitt:1962hi}, while $E_{ij}$ is related to the extrinsic curvature of equal-time hypersurfaces, see Appendix \ref{appA} for a detailed discussion. Furthermore, $\delta N$ and $\delta E_{ij}$ denote perturbations of the corresponding quantities over their background values. The `classical' theory \eqref{ggg} generates only the first two terms on the second line of \eqref{s_pi}, and all of the above analysis has assumed vanishing $\bar M$ and $\bar M'$ (as well as yet higher-derivative operators, implied by the ellipses). In practice, the latter coefficients are expected to be present, although suppressed in derivative expansion. In what follows, we assume nonzero $\bar M^2$ and $\bar M'^2$ in computing the quadratic action for $\zeta$ on extended genesis backgrounds \footnote{Both $\bar M^2$ and $\bar M'^2$ can in principle have either sign. The notation used for these coefficients only serves to emphasize their mass dimension. }. The results, given in \eqref{A'}-\eqref{C'} of Appendix \ref{appA}, are rather tedious and reluctant to simple analysis in their exact form. To simplify life, we will expand all relevant quantities to linear order in $\bar M^2$ and $\bar M'^2$, assuming these are small in the sense that higher order terms in the expansion give subleading corrections -- something we will justify \textit{a posteriori}. The procedure yields the following expressions for the kinetic coefficients\footnote{The signs are defined so that all kinetic coefficients have to be positive for complete stability (stability at all wavelengths) of the corresponding background.} $A$, $B$ and $C$ on backgrounds corresponding to the second, $\lambda\neq 0$ case of the previous section \begin{eqnarray} A&=&\frac{2}{3M_{\rm Pl}^2 H_0^4\lambda^2\tau^6}~\big(18 M_{\rm Pl}^4 H_0^4\lambda^2 \tau^6-3M_{\rm Pl}^2f^2H_0^2(5\lambda \tau-2)\tau^2+4 f^4\big )+p_1\bar M^2+p_2\bar M'^2,\nonumber \\ B&=&-\frac{2}{3}\frac{f^2}{H_0^2\lambda}\frac{1}{\tau^3}+p_3\bar M^2+p_4\bar M'^2+q_3\partial_t(\bar M^2)+q_4\partial_t(\bar M'^2),\nonumber \\ C&=&\frac{\bar M^2+\bar M'^2}{2\lambda^2},\nonumber \end{eqnarray} where we have defined $\tau \equiv t-t_*<0$ and introduced auxiliary coefficients $p_i$ and $q_i$, given as follows \begin{eqnarray} p_1&=&-\frac{1}{18 M_{\rm Pl}^8 H_0^8\lambda^4\tau^{12}}~\big (27 M_{\rm Pl}^4 H_0^4\lambda^2\tau^6-6M_{\rm Pl}^2 f^2 H_0^2(5\lambda\tau-2)\tau^2+8 f^4\big )^2, \nonumber \\ p_2&=&-\frac{1}{18 M_{\rm Pl}^8 H_0^8\lambda^4\tau^{12}}~\bigg [ 1107 M_{\rm Pl}^8 H_0^8\lambda^4\tau^{12}-1980M_{\rm Pl}^6 f^2H_0^6\lambda^3\tau^9 +792M_{\rm Pl}^6 f^2H_0^6\lambda^2\tau^8\nonumber \\&+&1428M_{\rm Pl}^4 f^4 H_0^4\lambda^2\tau^6-720 M_{\rm Pl}^4 f^4 H_0^4\lambda \tau^5+144M_{\rm Pl}^4 f^4 H_0^4\tau^4-480M_{\rm Pl}^2 f^6 H_0^2\lambda\tau^3\nonumber \\ &+&192 M_{\rm Pl}^2 f^6 H_0^2 \tau^2+64 f^8 \bigg ],\nonumber\\ p_3&=&-\frac{1}{18M_{\rm Pl}^6 H_0^6 \lambda^3\tau^9}~\bigg[81M_{\rm Pl}^6 H_0^6\lambda^3 \tau^9-18 M_{\rm Pl}^4 f^2 H_0^4 (8\lambda^2\tau^2-17\lambda\tau+8)\tau^4\nonumber \\ &+&84 M_{\rm Pl}^2 f^4 H_0^2 (\lambda \tau-2)\tau^2-16 f^6 \bigg],\nonumber \\ p_4&=&-\frac{1}{18M_{\rm Pl}^6 H_0^6 \lambda^3\tau^9}~\bigg[99M_{\rm Pl}^6 H_0^6\lambda^3 \tau^9-6M_{\rm Pl}^4 f^2 H_0^4 (26\lambda^2\tau^2-51\lambda\tau+24)\tau^4\nonumber \\ &+&84 M_{\rm Pl}^2 f^4 H_0^2 (\lambda \tau-2)\tau^2-16 f^6 \bigg],\nonumber \\ q_3&=&-\frac{1}{6 M_{\rm Pl}^4 H_0^4\lambda^3\tau^{6}}~\big[ 27M_{\rm Pl}^4 H_0^4 \lambda^2\tau^6-6M_{\rm Pl}^2 f^2 H_0^2(5\lambda \tau-2)\tau^2+8 f^4 \big],\nonumber \\ q_4&=&-\frac{1}{6 M_{\rm Pl}^4 H_0^4\lambda^3\tau^{6}}~\big[ 33M_{\rm Pl}^4 H_0^4 \lambda^2\tau^6-6M_{\rm Pl}^2 f^2 H_0^2(5\lambda \tau-2)\tau^2+8 f^4 \big]~.\nonumber \end{eqnarray} An immediate and important observation is that all of the coefficients $p_i$ and $q_i$ are sign-definite (negative) \textit{at all times}. Moreover, since different linear combinations of $\bar M^2$ and $\bar M'^2$ enter into $B$ and $C$, nothing prevents us from choosing the former pair of EFT coefficients such that both $B$ and $C$ are positive - thus avoiding any instability over the entire cosmological period of interest! Furthermore, we have found in the previous section that the speed of sound of the scalar mode tends to zero ($c_s^2\to 0^+$) in the asymptotic past for the backgrounds corresponding to EG. This suggests that $t\to -\infty$ is precisely the regime where higher-order corrections in the EFT for perturbations could play an important role. In fact, in the case that $\bar M, \bar M'$ fall off slower than $1/t^3$ at early times, higher-order effects give contributions that dominate over the leading-order piece in the coefficient $B$ at early times\footnote{This seemingly casts shadow on the very meaning of our expansion in small $\bar M, \bar M'$; fortunately, a closer inspection of \eqref{A'}-\eqref{C'} shows that the expansion parameters at $t\to -\infty$ are in fact $\bar M^2/M_{\rm Pl}^2$ and $\bar M'^2/M_{\rm Pl}^2$ -- meaning that next-order corrections in the series indeed give subleading effects.}. Focusing on the $t\to-\infty$ ghost condensate regime (and neglecting time derivatives of $\bar M, \bar M'$ for simplicity), we find \begin{eqnarray} \label{App} A&=&12 M_{\rm Pl}^2 +\mathcal{O}\(\bar M^2,\bar M'^2\)~, \\ \label{Bpp} B&=&-\frac{2}{3}\frac{f^2}{H_0^2\lambda}\frac{1}{t^3}-\frac{9}{2} \bar M^2-\frac{11}{2} \bar M'^2+\mathcal{O}\(\frac{\bar M^4}{M_{\rm Pl}^2},\frac{\bar M'^4}{M_{\rm Pl}^2}\)~, \\ \label{Cpp} C&=&\frac{\bar M^2+\bar M'^2}{2 \lambda^2}+\mathcal{O}\(\frac{\bar M^4}{M_{\rm Pl}^2\lambda^2},\frac{\bar M'^4}{M_{\rm Pl}^2\lambda^2}\) ~. \end{eqnarray} Again, since $B$ and $C$ involve different linear combinations of $\bar M^2$ and $\bar M'^2$, one can freely choose the values for the latter two coefficients, such that the theory is free from any instability\footnote{Note that there is in fact even more freedom: one could always make the coefficient $C$ positive by adding a term of the form $\sqrt{g_3} N R_3^{~2}$ (which does not affect the quadratic action for tensor perturbations) to \eqref{s_pi}, see the discussion in Appendix \ref{num}.}. Moreover, in the case that $\bar M, \bar M'$ drop off slower than $1/t^3$ for large and negative times, the asymptotic speed of sound of the scalar perturbation is set by the ratio $$c_s^2 = \frac{|9 \bar M^2+11\bar M'^2|}{24 M_{\rm Pl}^2}~,$$ and is not necessarily infinitesimally close to zero if at least one of the two EFT coefficients $\bar M$ and $\bar M'$ tends to a constant at early times\footnote{Note, that the squared speed of sound also sets the magnitude for the expansion parameter in \eqref{App}-\eqref{Cpp}.}. To summarize, we have found that beyond-the-leading-order structure of the effective theory for perturbations does possess enough freedom to allow to cure (weak) classical gradient instability. Whenever the speed of sound of the scalar mode vanishes at the leading order on the other hand, higher-order effects can push the corresponding solution into a completely stable direction. \section{Conclusions and future directions} \label{concl} Despite the extremely compelling picture of the early cosmology that standard slow-roll inflation provides us with, it is still fair to say that it is not the only possible one. The question of how far alternative scenarios can go in adequately describing the observed universe has been a strong motivation for expanding the theory space in non-standard directions. Perhaps the most dramatic departure from the inflationary paradigm corresponds to theories that violate the null energy condition, thereby allowing for a qualitatively different evolution of the early universe that, among other interesting features, is capable of smoothing out the Big-Bang singularity. That this can happen without instabilities for a universe starting out from the flat, Minkowski spacetime has been shown in Refs. \cite{Creminelli:2006xe,Creminelli:2010ba,Hinterbichler:2012fr}. A common feature of alternatives to inflation based on NEC violation is that they usually predict a strongly blue-tilted and unobservable (at least in the CMB experiments) spectrum of tensor perturbations. The ultimate reason for this lies in the fact that the phenomenologically interesting phase of cosmological evolution happens on quasi-flat backgrounds. Would then a detection of primordial $B$ modes in CMB polarization conclusively rule out these theories? In this paper we have argued that the answer to this question is negative. We have constructed explicit theories that lead to an early universe cosmology interpolating between a small/zero curvature maximally symmetric (dS or Minkowski) spacetime in the far past and an inflationary de Sitter spacetime, capable of generating a scale-invariant tensor spectrum of significant amplitude in the asymptotic future; this is possible because, at intermediate times, the system can \textit{strongly} violate the null energy condition ($\dot H\gg H^2$) as it happens in genesis models -- all without developing any kind of instability. The corresponding backgrounds can be viewed as a regularized extension of galilean genesis -- one for which none of the physical quantities grow beyond the cutoff scale. Alternatively, one can view them as a certain `UV' (or, to be more precise, an early-time) -complete realization of inflation, that leads to a (almost) flat pre-inflationary universe. Being deformations of the conformal galileon, the theories constructed above enjoy non-linearly realized emergent symmetries at both the early- and the late-time asymptotics. It is in fact precisely the nature of the asymptotics that determines the qualitative picture of the cosmological solutions of interest: these are described by quantum-mechanically stable, robust theories based solely on symmetry principles. An alternative, and very interesting realization of genesis cosmologies occurs in the context of the Dirac-Born-Infeld (DBI) models \cite{Hinterbichler:2012fr, Hinterbichler:2012yn}. Needless to say, it would be interesting to see how our construction carries over to theories enjoying asymptotic DBI-like symmetries. At this stage, the presented models are not intended as fully realistic, however upon slight adjustment they should become capable of facing observational challenges. While realistic model-building is beyond the scope of this paper, we briefly list the phenomenological questions that remain to be addressed. Above all, a mechanism for exiting the inflationary de Sitter regime/reheating has to be specified\footnote{This can be done \textit{e.g.} by giving a step function-like potential to the scalar as in ghost inflation -- a mechanism that can be implemented in a technically natural way since the asymptotic shift symmetry breaking becomes localized in field space in this case, see, \textit{e.g.} \cite{ArkaniHamed:2003uz}.}. Another question is that of the observed negative scalar tilt, which is not characteristic of NEC- violating inflationary theories with the inflaton being the field responsible for adiabatic perturbations. The negative tilt of density perturbations can arise from small shift-symmetry breaking effects (necessary to end the de Sitter phase), or through the standard mechanisms such as curvaton \cite{Enqvist:2001zp,Lyth:2001nq} or inhomogeneous reheating \cite{Dvali:2003em}. Furthermore, we have seen that on extended genesis backgrounds, the scalar perturbations are characterized by a relatively small speed of sound. Quite generally, small scalar speed of sound translates into large equilateral non-gaussianity \cite{Cheung:2007st} -- a result that follows solely from the requirement of nonlinearly realizing the broken time translation invariance\footnote{In addition, the small scalar speed of sound leads to an interesting effect, whereby the scalar perturbations of a given comoving wavelength freeze out earlier than the tensor modes leading to an enhancement of the tensor-to-scalar ratio, see \textit{e.g.} the recent discussion of Ref. \cite{Baumann:2014cja}. In our context, this could lead to a striking possibility of scalars freezing out in the genesis phase, while the tensors -- in the inflationary one. }. These, among other phenomenological aspects, will be discussed elsewhere. Putting aside phenomenology, our models serve as a proof of principle for the possibility to smoothly and stably connect an inflationary quasi-de Sitter universe to a much lower, or even zero-curvature, maximally symmetric spacetime in the asymptotic past -- all without exiting the regime of validity of the underlying EFT. \vskip 0.1 cm {\bf Acknowledgements}: It is a pleasure to thank Riccardo Barbieri, Paolo Creminelli, Gia Dvali, Gregory Gabadadze, Oriol Pujolas and Filippo Vernizzi for valuable discussions. The work of D.P. is supported in part by MIUR-FIRB grant RBFR12H1MW and by funds provided by Scuola Normale Superiore through the program "Progetti di Ricerca per Giovani Ricercatori". E.T. is supported in part by MIUR-FIRB grant RBFR12H1MW. The work of P.U. is supported in part by the DOE grant DE- SC0011784.
\section{Introduction}\setcounter{equation}{0} The spin-orbit force is a crucial ingredient in many parts of nuclear physics \cite{Greiner}. In the elementary shell model, nuclei are described as a collection of nucleons which do not directly interact. They only interact indirectly through an effective potential which gives rise to a one-particle Hamiltonian and consequently an energy spectrum. By the Pauli exclusion principle, levels of this energy spectrum are filled as the baryon number $B$ is increased. For special values of $B$, the spectrum hits a gap and the corresponding nucleus is tightly bound and very stable. These special values are called magic numbers and give rise to magic nuclei. The shell model works well near these. To obtain the correct magic numbers one must include a spin-orbit term in the single particle Hamiltonian \cite{Otto,Mayer}. This couples the spin of a nucleon to its orbital angular momentum $l$. The inclusion of this term breaks the degeneracy between states with the same value of $|l|$. States with spin and orbital angular momentum aligned are energetically favoured. For a nucleus with a few more nucleons than a magic number we can interpret its structure physically: a core made from a magic nucleus is surrounded by the other nucleons orbiting it. If we have one orbiting nucleon, its spin and orbital angular momentum are aligned in all but two cases, Antimony-$133$ and Bismuth-$209$. As more nucleons are added, other factors such as pairing make the interpretation more complicated. The spin-orbit force is strongest near the surface of the core and its physical meaning is lost within the core. Analogy with atomic physics points to an electromagnetic origin of the spin-orbit coupling but this turns out to have the wrong magnitude. The correct magnitude can be obtained by considering relativistic effects. They lead to a field theory where nucleons interact via mesons. The system can be solved approximately by neglecting quantum fluctuations of certain terms \cite{Ring}. While this technique is successful, it ignores the structure of nucleons and requires one to fit several parameters. Ideally these parameters would come from experiment but as the theory is phenomenological, effective masses and coupling constants must be used \cite{Ring2}. The spin-orbit force is also present in nucleon-nucleon interactions. It couples the orbital angular momentum to the sum of the spins of the nucleons, and can also be thought of as coming from meson interactions. The asymptotic form of the force has been successfully reproduced in the Skyrme model using a product ansatz which is valid only at large separations \cite{Riska} . In this paper we develop an idea in \cite{MSkySD} which provides an explanation for the spin-orbit force at shorter separations, inspired by the Skyrme model. We introduce the Skyrme model in section $2$, describing the important features which shape the spin-orbit interaction. We then set up a precise, simplified model of Skyrmion-Skyrmion interactions and solve it in section $3$, first for the simpler case of nucleon-nucleon interactions and then for the case of a nucleon interacting with a larger nucleus, which describes certain shell model configurations. \section{The Skyrme Model} The Skyrme model is a nonlinear field theory of pions which admits soliton solutions called Skyrmions \cite{Sk}. These are identified as nuclei with the topological charge $B$ of a Skyrmion equal to the baryon number of the system. A solution is generally represented by a surface of constant baryon density which is coloured to express the direction of the pion field as it varies over the surface; we use the same colouring scheme as in \cite{108}. The spherically symmetric $B=1$ solution, known as the hedgehog, is displayed in figure $1$. When two $B=1$ Skyrmions are widely separated we can approximate their interaction using an asymptotic expansion. One finds that among all configurations there is a special submanifold of maximal attraction between the Skyrmions called the attractive channel \cite{book}. This is easiest to interpret pictorially: in the attractive channel the separated Skyrmions have matching colours at the point of closest contact. Conversely, if the closest colours are opposite the Skyrmions repel. \begin{figure}[ht] \centering \includegraphics[width=4cm]{B1.png} \caption{The $B=1$ Skyrmion solution.} \label{fig:B1} \end{figure} Configurations tend to line up in the attractive channel in order to minimise potential energy. This concept remains useful for larger Skyrmions. As an example, consider the configuration in figure \ref{fig:B1B6}. Here, a $B=1$ Skyrmion is orbiting a $B=6$ Skyrmion. The system is shown in the attractive channel with red on both Skyrmions at their contact point. To stay in the attractive channel as it orbits, the $B=1$ Skyrmion must take a special orbital path. Specifically, it rolls around the equator of the larger solution completing three full rotations on its axis before returning to the initial position. The key observation is that the $B=1$ Skyrmions' orbital angular momentum is aligned with its spin. This is exactly what is required for the spin-orbit force in nuclei with $B$ one more than a magic number, except in the cases of Antimony-$133$ and Bismuth-$209$. It is the classical pion field structure of Skyrmions that provides the microscopic origin for the coupling. Many other Skyrmion pairs have paths like this which encourage spin-orbit coupling. The effect becomes stronger when the Skyrmions are closer together but loses meaning if they were to merge fully. This is consistent with the fact that the traditional spin-orbit force is strongest near the surface of the core nucleus. \begin{figure}[ht] \centering \includegraphics[width=8cm]{plotting-5.png} \caption{A $B=1$ Skyrmion close to a $B=6$ Skyrmion. The colours of closest contact are both red (unseen on the $B=1$ solution from this viewpoint) so the configuration is in the attractive channel.} \label{fig:B1B6} \end{figure} We will now try to work out the consequences of this classical spin-orbit coupling when the system is quantised. The usual procedure for quantising one Skyrmion is to use a rigid body approach to the classical minimal energy configuration, promoting its collective coordinates to quantum operators. Quantising the interaction between separated Skyrmions is more difficult and little progress has been possible using the full set of collective coordinates \cite{Manifolds}. Thus, we will only consider a toy model in two dimensions where we treat the Skyrmions as rigid discs. We will begin by carefully considering the simplest system possible: the interaction of two $B=1$ Skyrmions. \section{Discs interacting through a contact potential} \subsection{Two discs of equal size} Our model is based on taking $2$D slices of $3$D Skyrmion configurations, taking our inspiration from $B=1$ Skyrmion interactions. Figure \ref{fig:2Dslice}a shows separated $B=1$ solutions in the attractive channel. We can take a $2$D slice of this parallel to the $y$-$z$ plane and parallel to the $x$-$y$ plane to give us the systems in figures \ref{fig:2Dslice}b and \ref{fig:2Dslice}c. We now treat these $2$D objects as rigid discs, at fixed separation, interacting through a potential which depends only on their colouring. \begin{figure}[ht] \centering \includegraphics[width=15cm]{2Dslices.pdf} \caption{(a) Two $B=1$ Skyrmions in the attractive channel. (b) The rolling configuration. (c) The sliding configuration} \label{fig:2Dslice} \end{figure} To remain in the attractive channel the discs in figures \ref{fig:2Dslice}b and \ref{fig:2Dslice}c must, respectively, roll and slide around each other. For now, we will consider the rolling configuration. Labelling the discs as $1$ and $2$ we introduce the angular coordinates as in figure \ref{fig:angles}. The angles $\alpha_1$ and $\alpha_2$ represent the orientation of the discs with respect to their own axes. These are measured anti-clockwise and are zero when white points up, as in figure \ref{fig:2Dslice}b. The coordinate $\beta$ labels the orbital orientation of the discs while $r$ is the (fixed) distance between the disc centres. \begin{figure}[ht] \centering \includegraphics[width=6cm]{angles.pdf} \caption{The angles $\alpha_1$, $\alpha_2$ and $\beta$.} \label{fig:angles} \end{figure} Each of the coordinates has range $2\pi$ but $\beta \to \beta + \pi$ also returns the system to the attractive channel. As such, the potential must be periodic under full rotations of either disc and under half an orbital rotation. It should also only depend on the colouring: the simplest choice is a cosine potential. Thus, a classical Lagrangian which describes the system is \begin{equation} L = \frac{1}{2}I_1\dot{\alpha}_1^2 + \frac{1}{2}I_2\dot{\alpha}_2^2 + \frac{1}{2}\mu r^2\dot{\beta}^2+ k\cos\left(2\beta - \alpha_1 - \alpha_2\right) \end{equation} where $I_1$, $I_2$ are the disc moments of inertia, $\mu$ is the reduced mass of the system and $k>0$ is the strength of the potential. The argument of the potential measures the difference in colour at the closest points. The discs are identical so $I_1=I_2=I$ and we may write $\mu r^2$ in terms of $I$ by introducing a dimensionless separation parameter $d$ and setting $\mu r^2 = 4d^2I$. This simplifies the Lagrangian to \begin{equation} \label{lag} L = \frac{1}{2}I\left(\dot{\alpha}_1^2 + \dot{\alpha}_2^2 + 4d^2\dot{\beta}^2\right)+ k\cos\left(2\beta - \alpha_1 - \alpha_2\right) \,. \end{equation} Classically, the lowest energy solution satisfies $2\beta - \alpha_1 - \alpha_2 = 0$. This forces the discs into the attractive channel as if they were cogwheels; the first cog rolls around the second, fixed cog. If they stay in the attractive channel for all time, we can differentiate this condition to obtain a relation between velocities: $2\dot{\beta} - \dot{\alpha}_1 - \dot{\alpha}_2 = 0$. Introducing the classical conjugate momenta to the coordinates \begin{equation} s_1 = I\dot{\alpha}_1\,,\, s_2 = I\dot{\alpha}_2 \text{ and } l = 4d^2I\dot{\beta}\,, \end{equation} we can rewrite the above velocity relation as \begin{equation} \label{cog} l - 2d^2(s_1+s_2) = 0\,. \end{equation} Later we will see that this combination of spins and angular momentum has an important role to play in the quantum picture too. The Lagrangian \eqref{lag} has two linearly independent continuous symmetries. The first corresponds to all angles increasing by the same amount. This leads to conservation of total angular momentum \begin{equation} \label{Jspin} \mathcal{J} = I\left(\dot{\alpha}_1+\dot{\alpha}_2+4d^2\dot{\beta}\right) = s_1 + s_2 + l\,. \end{equation} The other conserved quantity is generated by one disc spinning at the same speed as the other but in the opposite direction. Since this quantity can be interpreted purely in terms of the colour fields moving, we label it as the total isospin in analogy with the full Skyrme model. It has the form \begin{equation} \label{Ispin} \mathcal{I} = I\left(\dot{\alpha}_1-\dot{\alpha}_2\right) = s_1-s_2\,. \end{equation} We can take advantage of these symmetries by changing coordinates and reducing the problem's degrees of freedom from three to one. Before doing this, we should consider the discrete symmetries of the system which occur since the configuration space is a $3$-torus. First let us solve the problem for $k=0$ where the Hamiltonian becomes that of a free particle on a $3$-torus. After canonical quantisation, the Hamiltonian has the form \begin{equation} \hat{\mathcal{H}} = -\frac{1}{2I}\left( \frac{\partial^2}{\partial\alpha_1^2}+\frac{\partial^2}{\partial\alpha_2^2}+\frac{1}{4d^2}\frac{\partial^2}{\partial\beta^2} \right) \end{equation} where we have set $\hbar = 1$. The wavefunction has the form \begin{equation} \label{sss} \psi_{\text{free}}\left(\alpha_1,\alpha_2,\beta\right) = e^{i(s_1\alpha_1 + s_2\alpha_2 + l\beta)} \end{equation} with corresponding energy \begin{equation} E_{\text{free}} = \frac{1}{2I}\left( s_1^2 + s_2^2 + \frac{1}{4d^2}l^2 \right)\,. \end{equation} The quantities $s_1$, $s_2$ and $l$ are the quantum numbers corresponding to the spins and orbital angular momentum of the free discs. As we are modelling Skyrmions, the discs are treated as fermions. Thus, the wavefunction picks up a minus sign under full disc rotations: $\alpha_1 \to \alpha_1 + 2\pi$ and $\alpha_2 \to \alpha_2 + 2\pi$. This means $s_1$ and $s_2$ are both half-integers. The system is also invariant under $\beta \to \beta + 2\pi$ and as such $l$ must be an integer. While these quantities do not remain good quantum numbers when the potential is turned on, they do remain important due to Bloch's theorem. This says that there exists a basis of energy eigenstates of the form \begin{equation} \label{Bloch} \psi\left(\alpha_1,\alpha_2,\beta\right) = e^{i(s_1\alpha_1 + s_2\alpha_2 + l\beta)}u\left(\alpha_1,\alpha_2,\beta\right)\,, \end{equation} where $u$ is periodic on the $3$-torus, and has the same periodicity as the potential. This theorem is generally used in an infinite lattice but we are on a torus. As such $s_1$, $s_2$ and $l$ have discrete allowed values instead of continuous ones. They are also usually defined up to a vector in the reciprocal lattice, a discrete lattice in $3$D. However we fix their value by insisting that \begin{equation} u\left(\alpha_1, \alpha_2, \beta\right)|_{k=0} \equiv 1\,. \end{equation} There is one state per cell in the reciprocal lattice. Thus we can understand $s_1$, $s_2$ and $l$ as labelling a particular lattice cell. We will see later that energy states from different cells do not cross when the potential is turned on and as such these labels are good for tracking the energy states as $k$ increases. To make progress we must now change coordinates to take advantage of the continuous symmetries from earlier. We introduce new coordinates $(\gamma,\xi,\eta)$. Two of these should give rise to the conjugate momenta corresponding to $\mathcal{J}$ and $\mathcal{I}$. That is \begin{align} -i\frac{\partial}{\partial\xi} &= -i\left(\frac{\partial}{\partial\alpha_1} + \frac{\partial}{\partial\alpha_2}+\frac{\partial}{\partial\beta}\right)\\ -i\frac{\partial}{\partial\eta} &= -i\left(\frac{\partial}{\partial\alpha_1} - \frac{\partial}{\partial\alpha_2}\right) \,. \end{align} Note that these operators commute with the potential in \eqref{lag}. We may define $\gamma$ to be the coordinate in the potential. If we also insist on a diagonal quadratic kinetic term in the Hamiltonian we arrive at a unique coordinate transformation \begin{equation} \label{changeo} \begin{pmatrix} \gamma \\ \xi \\ \eta \end{pmatrix} = \begin{pmatrix} -1 & -1 & 2 \\ \frac{1}{2+4d^2} & \frac{1}{2 + 4d^2} & \frac{4d^2}{2+4d^2} \\ \frac{1}{2} & -\frac{1}{2} & 0 \end{pmatrix} \begin{pmatrix} \alpha_1 \\ \alpha_2 \\ \beta \end{pmatrix} \,. \end{equation} This transforms the Hamiltonian to \begin{equation} \hat{\mathcal{H}} = -\frac{1}{2I}\left(\frac{1+2d^2}{d^2}\frac{\partial^2}{\partial\gamma^2} + \frac{1}{2+4d^2}\frac{\partial^2}{\partial\xi^2} + \frac{1}{2}\frac{\partial^2}{\partial\eta^2}\right) - k\cos\gamma\,. \end{equation} Since the $\xi$ and $\eta$ contributions are purely kinetic, the wavefunction has the form \begin{equation} \psi(\gamma,\eta,\xi) = e^{i\mathcal{J}\xi}e^{i\mathcal{I}\eta}\chi(\gamma)\,. \end{equation} Moreover, after applying the coordinate transformation, comparison with \eqref{sss} and \eqref{Bloch} tells us that \begin{align} \mathcal{J} &= s_1 + s_2 + l\,, \\ \mathcal{I} &= s_1 - s_2 \end{align} as in the classical equations \eqref{Jspin} and \eqref{Ispin}, and that \begin{equation} \chi(\gamma) = e^{iq_\gamma \gamma}\tilde{u}(\gamma)\,, \end{equation} where \begin{equation} q_\gamma = \frac{1}{2+4d^2}\left(l - 2d^2(s_1 + s_2)\right) \end{equation} and $\tilde{u}(\gamma)$ has period $2\pi$. Once again, we fix $q_\gamma$ so that $\tilde{u}|_{k=0} \equiv 1$. From earlier, we find that $\mathcal{I}$ and $\mathcal{J}$ can take any integer values. The free system now has the wavefunction \begin{equation} \psi_{\text{free}}(\gamma,\eta,\xi) = e^{i\mathcal{J}\xi}e^{i\mathcal{I}\eta}e^{iq_\gamma\gamma} \end{equation} with corresponding energy \begin{equation} E_{\text{free}} = \frac{1+2d^2}{2d^2I}q_\gamma^2 + \frac{1}{(4 + 8d^2)I}\mathcal{J}^2 + \frac{1}{4I}\mathcal{I}^2\,. \end{equation} For fixed $\mathcal{J}$ and $\mathcal{I}$, the allowed values of $q_\gamma$ are separated by integers, though the fractional part of $q_\gamma$ depends on $d$ and $\mathcal{J}$. Combining everything, the problem reduces to the Schr\"{o}dinger equation \begin{align} \label{schro} -\frac{1+2d^2}{2d^2I}\frac{d^2 }{d \gamma^2}\left( e^{iq_\gamma \gamma}\tilde{u}\right) - k\cos\gamma \, e^{iq_\gamma \gamma}\tilde{u} &= \left( E - \frac{\mathcal{J}^2}{(4+8d^2)I} - \frac{\mathcal{I}^2}{4I} \right) e^{iq_\gamma \gamma}\tilde{u}\\ &\equiv E_\gamma e^{iq_\gamma \gamma}\tilde{u}\,. \end{align} This is the Mathieu equation, which has been extensively studied \cite{mathi}. We will now consider it with our physical picture in mind. The energy has separated into two parts -- one depends on $\mathcal{J}$ and $\mathcal{I}$ and has no $k$ dependence. The other only depends on the $\gamma$ sector. The potential does not mix states with different $\mathcal{I}$ and $\mathcal{J}$. Thus, we can fix these values and focus on calculating $E_\gamma$. We can understand the system when $k$ is small by using perturbation theory. Note that the dimensionless small quantity is really $kI$. The energy, to second order in $kI$ is \begin{equation} \label{perteng} E_{\gamma,\text{pert}} = \frac{1+2d^2}{2d^2I}q_\gamma^2 + (kI)^2\frac{d^2}{(1+2d^2)I}\frac{1}{4q_\gamma^2 -1}\,. \end{equation} The most important thing to note is that for fixed $\mathcal{J}$ and $\mathcal{I}$, since the allowed values of $q_\gamma$ are separated by integers, there is a unique state which satisfies $4q_\gamma^2 - 1 < 0$. Thus there is one state whose energy decreases after perturbation, with $|q_\gamma| \leq \frac{1}{2}$. We call states which satisfy this condition energetically favourable. At $|q_\gamma| = \frac{1}{2}$ equation \eqref{perteng} breaks down and degenerate perturbation theory must be used. It tells us that the energy spectrum develops a gap at each of these points leaving a separated energy band for $|q_\gamma| < \frac{1}{2}$ which does not touch the rest of the spectrum. The other degenerate points ($|q_\gamma| = 1, \frac{3}{2}, 2$...) lead to singularities in the perturbative energy spectrum at higher orders. Degenerate perturbation theory tells us, once again, that a gap occurs at each of these points. Thus after perturbation we are left with an energy spectrum divided into non-touching bands as seen in figure \ref{fig:pereng}. Degenerate points are identified and as such each band is an integer long. For example, one of the bands is $q_\gamma \in [-1,-\frac{1}{2}] \cup [\frac{1}{2},1]$. Since the allowed values of $q_\gamma$ are separated by an integer there is exactly one state per band. This explains why $q_\gamma$ is a good label: due to the gaps in the spectrum we can follow a free state as $k$ increases without having to worry about crossing except at degenerate points. Even there, the uncertainty is only between two states and most degeneracies only occur for special values of $d$. As such, we won't consider them carefully. \begin{figure}[ht] \centering \includegraphics[width=12cm]{perteng.pdf} \caption{How the energy spectrum changes after perturbation. $E_{\text{free}}$ is the spectrum for $k=0$; $E_{\text{pert}}$ is the spectrum for small $kI$. The dots represent an example of an allowed value of $q_\gamma$. In this case we take $(\mathcal{I},\mathcal{J}) = (0,1)$ and $d = 1$ which gives $q_\gamma \equiv \frac{1}{3} (\text{mod } 1)$. Note that there is one allowed state per separated band.} \label{fig:pereng} \end{figure} For large $k$ we may use a tight binding (\text{tb}) limit. This approximation relies on the wavefunction being concentrated within each unit cell in $\gamma$ with negligible overlap. Then the total wavefunction can be written as a sum of isolated wavefunctions which solve Schr\"odinger's equation within the unit cell. These isolated wavefunctions must be the same at each site due to the periodicity of $\tilde{u}$. Bloch's theorem allows for the total wavefunction to pick up a phase between cells meaning the solution of \eqref{schro} is of the form \begin{equation} e^{iq_\gamma\gamma}\tilde{u}_{\text{tb}}(\gamma) = \sum_{m \in \mathbb{Z}} \phi(\gamma - 2\pi m) e^{2\pi iLm} \end{equation} where $\phi$ is the isolated wavefunction and $L$ is some constant. The periodicity of $\tilde{u}$ fixes $L$ to be $q_\gamma$. Thus our total, tight binding wavefunction is \begin{equation} \psi_{\text{tb}}(\gamma, \xi, \eta) = e^{i\mathcal{J}\xi}e^{i\mathcal{I}\eta}\sum_{m\in \mathbb{Z}}\phi(\gamma - 2\pi m)e^{2\pi i mq_\gamma}\,. \end{equation} We are left to find $\phi$. Since $k$ is large, we assume that the wavefunction is concentrated near the minimum of the potential. We can expand the potential near this point, which gives \begin{equation} \label{pot} -k\cos\gamma \approx -k\left(1 - \frac{\gamma^2}{2} + \frac{\gamma^4}{4!}\right) \end{equation} and reduces the Schr\"odinger equation \eqref{schro} to \begin{equation} \label{phischro} -\frac{1+2d^2}{2d^2I}\frac{d^2\phi}{d\gamma^2} - k\left(1 - \frac{\gamma^2}{2} + \frac{\gamma^4}{24}\right)\phi = E_\gamma\phi\,, \end{equation} where $kI$ is large. If we temporarily ignore the $\gamma^4$ term in the potential then this truncated Schr\"odinger equation is just a simple harmonic oscillator which can be solved by standard methods. The non-normalised eigenstates are given by \begin{equation} \phi_N(\gamma) = H_N\left(\left(\frac{kd^2I}{1+2d^2}\right)^\frac{1}{4}\gamma\right)\exp\left(-\left(\frac{kd^2I}{4+8d^2}\right)^\frac{1}{2}\gamma^2\right) \end{equation} where $H_N$ are the Hermite polynomials. We can then use these to find the energies to $O(1)$ in $k$. They are \begin{equation} \label{endo} E_{\gamma,n} = -k + \sqrt{k}\sqrt{\frac{1+2d^2}{d^2I}}\left(N+\frac{1}{2}\right) - \frac{1+2d^2}{32d^2I}\left(2N^2 + 2N + 1\right) + O\left( \frac{1}{ \sqrt{k} } \right)\,. \end{equation} The $O(k)$ term is from the constant in the potential. The $O(\sqrt{k})$ term is the usual harmonic oscillator energy, and the $O(1)$ term is the contribution from the $\gamma^4$ term in the potential, evaluated by first order perturbation theory. We have ignored all overlap terms between cells, but these are exponentially suppressed for large enough $k$. Due to the lattice structure, the labels we used for the free states continue to label the states in the tight binding limit. Since there is no crossing for fixed $\mathcal{I}$ and $\mathcal{J}$, the $N^\text{th}$ excited free state (which has the label $q_\gamma$ where $q_\gamma \in [-\frac{N+1}{2},-\frac{N}{2}]\cup[\frac{N}{2},\frac{N+1}{2}]$) flows smoothly to the state labelled by $N$ in the tight binding limit. This is confirmed by numerical calculations as seen in figure \ref{fig:engnums}, which shows the analytic and numerical energies as a function of $k$ for the four lowest energy states for fixed $(\mathcal{I}, \mathcal{J}) = (0,1)$. The eigenvalues $E_\gamma$ are found using a shooting method. \begin{figure}[ht] \centering \includegraphics[width=13.8cm]{energiesnumerics.pdf} \caption{The energy spectrum for $(\mathcal{I},\mathcal{J}) = (0,1)$ and $d=1$ as $k$ varies. As in figure \ref{fig:pereng}, these values give $q_\gamma \equiv \frac{1}{3} (\text{mod } 1)$. Our analytic expressions are represented by the bold lines while numerical results are displayed as dots. The $N^\text{th}$ excited free state (and thus the free state in the $N^\text{th}$ band) flows to the $N^\text{th}$ excited state of the tight binding limit.} \label{fig:engnums} \end{figure} From the numerical data in figure $6$ we see that the analytic expressions \eqref{perteng} and \eqref{endo} have different regions of validity depending on which state we examine. We can explain this as follows. The large $k$ calculation relied on two approximations: that the wavefunction is concentrated within a unit cell and that it is concentrated within a region where we may expand the potential to quartic order. If we satisfy the second constraint we certainly satisfy the first so we shall examine the second. The expansion \eqref{pot} is, very roughly, good for $|\gamma| < 2$. Thus, we need the wavefunction to be decaying exponentially there. For large $\gamma$, $\phi_n$ is of the form \begin{align} \phi_N &\sim \gamma^N\kappa^\frac{N}{4}\exp\left(-\frac{\gamma^2}{2}\kappa^\frac{1}{2}\right) \nonumber \\ &= \exp\left( N\log\gamma + \frac{N}{4}\log\kappa - \frac{\gamma^2\kappa^{\frac{1}{2}}}{2}\right) \end{align} where we have defined $\kappa = \frac{kd^2I}{1+2d^2}$. For the wavefunction to be concentrated within $-2 < \gamma < 2$ we require \begin{equation} N\log2 + \frac{N}{4}\log\kappa - 2\kappa^{\frac{1}{2}} < -c \end{equation} where $c$ is some positive constant. We see that as $N$ increases we need a larger $\kappa$, and hence $kI$, for our approximation to be valid, as the numerical results confirm. The regions of validity of the small $k$ perturbative energy expansion \eqref{perteng} can be explained by calculating the next non-trivial term. It is \begin{equation} k^4\frac{(d^2I)^3}{(1+2d^2)^3}\frac{20q_\gamma^2 + 7}{(4q_\gamma^2-1)^3(4q_\gamma^2-4)} \,. \end{equation} Away from degenerate points, this goes as $k^4q_\gamma^{-6} $ and as such is small for states with large $q_\gamma$. This explains why the perturbative energy calculation works for a larger range of $k$ for states with larger $q_\gamma$. The problem has now been solved in both small and large $k$ limits. Thanks to the lattice structure we can extrapolate the free states to the large $k$ states without fear of crossing between states. We have found that the energetically favourable states in the large $k$ limit come from free states which satisfy \begin{equation} \label{ineq} |q_\gamma| = \frac{1}{2+4d^2}\left| l - 2d^2\left(s_1+s_2\right)\right)| \leq \frac{1}{2}\,. \end{equation} This is our form of spin-orbit coupling. States with orbital angular momentum and spins aligned are more likely to satisfy the inequality while they are less likely to if they are anti-aligned. Note the connection between the classical minimum energy condition \eqref{cog} and our energetically favourable state condition \eqref{ineq}. To be more definite let us fix the spins of the discs be $\pm\frac{1}{2}$. Then $s_1 + s_2$ can be $1,0$ or $-1$. Take $s_1+s_2=1$ first. Energetically favoured states satisfy \begin{equation} |l - 2d^2| \leq 1+2d^2\,. \end{equation} Thus these states have orbital angular momentum $l \in [-1,1+4d^2]$. As $l$ is usually positive, spin and orbital angular momentum are usually aligned. The extreme case, $l = -1$, corresponds to a degenerate point in the energy spectrum. Here, our labels lose meaning and we cannot distinguish between the free states $(s_1,s_2,l) = (\frac{1}{2},\frac{1}{2},-1)$ and $(s_1,s_2,l) = (-\frac{1}{2},-\frac{1}{2},1)$ as the potential is turned on. Both of these have spin and orbital angular momentum anti-aligned. The result is essentially the same for $s_1+s_2 = -1$. Here $l$ is always non-positive, except for the degenerate state. As this example demonstrates, the direction of the spins is correlated with the direction of the orbital angular momentum for most of the energetically favoured states. When $s_1+s_2 = 0$ the condition \eqref{ineq} reduces to \begin{equation} |l| \leq 1+2d^2\,. \end{equation} This time there is no spin-orbit coupling as the orbital angular momentum has no preferred direction. Figure \ref{fig:fullspec} displays how the energy spectrum changes as $k$ is turned on for the lowest energy states which satisfy $\mathcal{I} = 0$ and $s_1 = \frac{1}{2}$. We focus on these states as this is where the spin-orbit force is present in our model. Note that states with equal $|l|$ in the free case become non-degenerate for positive $k$, just as they do in traditional spin-orbit coupling. For this figure, we take $I = 1$, $d=0.9$, with $d$ not equal to $1$ so that we avoid certain degeneracies. \begin{figure}[ht] \centering \includegraphics[width=14cm]{fullspecs.pdf} \caption{The energy spectrum for some low lying states with various values of $\mathcal{J}$, with $\mathcal{I} = 0$ and $s_1 = \frac{1}{2}$. Each is labelled by their $(s_1,s_2,l;\mathcal{J})$ value at $k=0$. In all but the extreme case, $l=-1$, the energetically favoured states have spin and orbital angular momentum aligned.} \label{fig:fullspec} \end{figure} In the large $k$ limit only those states which came from free states with $|q_\gamma| \leq \frac{1}{2}$ are contenders for the ground state. These are then ordered by the $\mathcal{I},\mathcal{J}$ energy contribution. This limit is exactly rigid body quantisation and in the strict limit the wavefunction is a delta function, the system completely fixed in the attractive channel. Physically, we expect the true strength of $k$ to be between the two limits we understand analytically. This is also seen in the traditional spin-orbit force: the coupling is strong enough that it has an effect on the energy spectrum but weak enough that an understanding of the spectrum without the force is vital too. We may do an analogous calculation for the sliding configuration from figure \ref{fig:2Dslice}c. The calculation is very similar to the one above and the main physical consequence is that the energetically favoured states come from free states with small $s_1 - s_2$. Thus, the sliding configuration couples the spins. This is what is required for the tensor force -- another key ingredient in nucleon-nucleon interactions. Thus, our model unifies the spin-orbit force and the tensor force while giving them both a classical microscopic origin. In the full $3$D model both sliding and rolling motion can occur simultaneously and both need to be taken into account at the same time. \subsection{Unequal discs} Consider a generalisation of the system. Now a small disc orbits a larger one as seen in figure \ref{fig:lilbig}, with small and large discs labelled $1$ and $2$ respectively. Let the colour field repeat $n$ times along the edge of the large disc. This is a model for a nucleus with baryon number one more than a magic number, with a single nucleon orbiting a core. The core is generally a boson and this is how we treat the large disc. Defining our variables analogously to the variables in the previous section and using the initial configuration as in figure \ref{fig:lilbig}, the Lagrangian \eqref{lag} is modified to \begin{equation} L = \frac{1}{2}I_1\dot{\alpha}_1^2 + \frac{1}{2}I_2\dot{\alpha}_2^2 + \frac{1}{2}\mu r^2\dot{\beta}^2+ k\cos\left((n+1)\beta - \alpha_1 - n\alpha_2\right)\,. \end{equation} \begin{figure}[ht] \centering \includegraphics[width=9cm]{biglil.pdf} \caption{A small disc orbiting a large disc.} \label{fig:lilbig} \end{figure} The classical conserved quantities are now \begin{align} \mathcal{J} &= I_1\dot{\alpha_1} + I_2\dot{\alpha_2} + \mu r^2\dot{\beta} = s_1+s_2+l\,, \\ \mathcal{I} &= nI_1\dot{\alpha_1} - I_2\dot{\alpha_2}=ns_1 - s_2 \,. \end{align} We are using the same notation as before: $l$ is the orbital angular momentum while $s_i$ is the spin of disc $i$. The classical minimum energy solution is when the discs are locked in the attractive channel and thus act like cogwheels. This gives a condition on the momenta of the system as follows: \begin{equation} \label{classact} I_1I_2(n+1)l - I_2\mu r^2s_1 - I_1\mu r^2ns_2 = 0\,. \end{equation} We can change coordinates so that the potential depends on one angle, $\gamma$, while the others are conjugate to $\mathcal{J}$ and $\mathcal{I}$. Further, we can insist that the Hamiltonian splits into two independent sectors (one depending only on $\mathcal{J}$ and $\mathcal{I}$, the other determined purely by the $\gamma$ sector) as we did in the previous section. Once again, this gives a unique coordinate transformation \begin{equation} \label{changeo2} \begin{pmatrix} \gamma \\ \xi \\ \eta \end{pmatrix} = \begin{pmatrix} -1 & -n & n+1 \\ \frac{I_1I_2(n+1)}{C} & \frac{I_1I_2n(n+1)}{C} & \frac{\mu r^2(I_2+I_1n^2)}{C} \\ \frac{I_1(I_2 + I_2n + \mu r^2n)}{C} & -\frac{I_2(I_1+I_1n+\mu r^2)}{C} & \frac{\mu r^2(I_2-I_1n)}{C} \end{pmatrix} \begin{pmatrix} \alpha_1 \\ \alpha_2 \\ \beta \end{pmatrix} \end{equation} where $C = I_1I_2(n+1)^2 + I_1\mu r^2 n^2+ I_2\mu r^2$. This, combined with Bloch's theorem gives us the form of the wavefunction after canonical quantisation. It is \begin{equation} \psi(\gamma,\eta,\xi) = e^{i\mathcal{J}\xi}e^{i\mathcal{I}\eta}e^{iq_\gamma\gamma}\tilde{w}(\gamma)\,. \end{equation} where $\tilde{w}$ has period $2\pi$ and $q_\gamma = (I_1I_2(n+1)l - I_2\mu r^2s_1 - I_1\mu r^2ns_2)C^{-1} $. Since the small disc is a fermion, $s_1$ must be a half-integer while $s_2$ and $l$ are both integers. Once again, the allowed values of $q_\gamma$ are separated by an integer. The Schr\"odinger equation is now \begin{equation} -\frac{C}{2I_1I_2\mu r^2}\frac{d^2}{d\gamma^2}\left(e^{iq_\gamma\gamma}\tilde{w}\right) - k\cos\gamma \, e^{iq_\gamma\gamma}\tilde{w} = E_\gamma e^{iq_\gamma\gamma}\tilde{w}\,, \end{equation} where the energy of the system is \begin{align} E &= E_\gamma + \frac{I_1n^2 + I_2}{2C}\mathcal{J}^2 + \frac{I_2-I_1n}{C}\mathcal{I}\mathcal{J} + \frac{I_1+I_2+\mu r^2}{2C}\mathcal{I}^2\\ &\equiv E_\gamma + E_{\mathcal{I},\mathcal{J}}\,. \end{align} This is simply equation \eqref{schro} with an adjusted mass. Thus we may apply all our analysis from the previous section to this problem; namely we can reuse the equations \eqref{perteng} to \eqref{endo} with the replacement \begin{equation} \frac{1+2d^2}{d^2I} \to \frac{C}{I_1I_2\mu r^2}\,. \end{equation} The physical consequence is that when $k$ is increased, the energetically favourable states have small \begin{equation} q_\gamma = \frac{1}{ I_1I_2(n+1)^2 + I_1\mu r^2 n^2+ I_2\mu r^2}\left( I_1I_2(n+1)l - I_2\mu r^2s_1 - I_1\mu r^2ns_2 \right)\,. \end{equation} Note the relationship between this and the classical condition \eqref{classact}. To gain more insight we must estimate the moments of inertia. First we assume that the circumference of the large disc is $n$ times the circumference of the small one. Then we use the Skyrmion inspired approximation that the radius of solutions with baryon number $B$ scales as $B^{\frac{1}{3}}$ and that their mass scales linearly with $B$. Finally, we assume that the discs are touching. These give us $I_2$ and $\mu r^2$ in terms of $I_1 $ as follows: \begin{equation} I_2 = n^5I_1 \,, \quad \mu r^2 = \frac{2n^3}{n^3 + 1}(n+1)^2I_1\,. \end{equation} It follows that \begin{equation} \frac{C}{I_1I_2\mu r^2} = \frac{3}{2I_1}\,\frac{n^3+1}{n^3} \end{equation} and \begin{equation} q_\gamma = \frac{1}{3(n+1)}\left( l - \frac{2n^3}{n^2-n+1}s_1 - \frac{2}{n(n^2-n+1)}s_2\right)\,. \end{equation} In \eqref{classact} we saw that the classical minimum energy solution obeyed $q_\gamma = 0$. If we also demand that the core is inert ($s_2=0$) then $l$ scales as $2ns_1$ for large $n$. This gives a natural explanation why orbital angular momentum increases as the size of the core increases, a relationship obeyed by the first few magic nuclei. We also see that if $s_2$ is non-zero, its contribution does not have much effect on the value of $q_\gamma$; the most important contribution is from the first two terms. Naively this looks promising: after quantisation, energetically favoured states obey $|q_\gamma| \leq \frac{1}{2}$ and this can be achieved by having $s_1$ and $l$ aligned. However, the number of energetically favoured states is rather large. To be concrete, let us fix $s_1 = \frac{1}{2}$ and $s_2 = 0$ from now on. Then \begin{equation} q_\gamma = \frac{1}{3(n+1)}\left(l - \frac{n^3}{n^2-n+1}\right)\,. \end{equation} To satisfy $|q_\gamma| \leq \frac{1}{2}$ we require \begin{equation} \label{lrange} l \in \left[ -\frac{n^3+3}{2(n^2-n+1)},\frac{5n^3+3}{2(n^2-n+1)} \right] \,. \end{equation} Thus, the restriction to energetically favourable states is in fact not very limiting and the range of allowed values of $l$ grows with $n$. The centre of this range corresponds to the classical minimum energy solution, $q_\gamma = 0$. In the $k=0$ limit the states are ordered by $|l|$. As $k$ increases we become more interested in the energetically favoured states. These are ordered, in the extreme large $k$ limit, by $E_{\mathcal{I},\mathcal{J}}$. In terms of $l$ this quantity is \begin{align} E_{\mathcal{I},\mathcal{J}} &= \frac{1}{24I_1n^3(n^3+1)}\left( 4l^2\left(n^2-n+1\right)^2 + 4ln^3\left(n^2-n+1\right) + n^3\left(n^3+3\right)\right) \nonumber \\ &= \frac{1}{24I_1n^3(n^3+1)}\left(4(n^2-n+1)^2\left(l + \frac{n^3}{2(n^2-n+1)}\right)^2 + 3n^3\right)\,. \end{align} This means that the states are ordered energetically by the magnitude of $|l +\frac{n^3}{2(n^2-n+1)}|$. From comparison with \eqref{lrange} we see that the state with minimal $E_{\mathcal{I},\mathcal{J}}$ lies within the energetically favoured range of $l$ values. Thus the ground state of the system in the large $k$ limit has spin and orbital angular momentum anti-aligned as $l$ is negative, going against our classical intuition. \begin{figure}[hr] \centering \includegraphics[width=14cm]{numsbiglilhi.pdf} \caption{Energy for a variety of low lying states of unequal discs with $n=3$, as a function of $k$. Here all states with $s_1 =\frac{1}{2}$, $s_2 = 0$ and $l \in [-4,4]$ are shown. For large $|l|$ the states with $s_1$ and $l$ aligned are favoured. However for small $|l|$, the opposite is true.} \label{fig:lilbig2} \end{figure} Let us consider $n=3$ in detail to illustrate these points more concretely. Here, there are twelve energetically favoured states, with $l \in [-2,9]$. Two of these have $l$ and $s_1$ anti-aligned and these two are the lowest energy states in the large $k$ limit. However, most of the energetically favoured states do have spin and orbital angular momentum aligned. The energy, as a function of $k$, of the states with $l \in [-4,4]$ is plotted in figure \ref{fig:lilbig2}. \pagebreak \section{Conclusions} The Skyrme model provides a classical microscopic origin for the spin-orbit force based on the classical pion field structure. In this paper, we have constructed a model of interacting Skyrmions based on discs interacting through a contact potential which depends only on their relative colouring. The classical behaviour resembles a pair of cogwheels and our quantisation of the model has shown that most low energy states have their spin and orbital angular momentum aligned. However, the ground state does not. To make any real predictions from the model we must extend it to three dimensions. This is considerably more difficult as there will be three relative orientations on which the potential depends, instead of one. There is also work to be done in the Skyrme model itself. Dynamical solutions of the model which look like a $B=1$ Skyrmion orbiting a core have not yet been found. \section*{Acknowledgements} C. J. Halcrow is supported by an STFC studentship.
\section{Introduction}\label{sec:intro} \par It is expected that in the early, metal-poor Universe the formation of massive stars was favored. These stars may have played an important role in the reionization of the gas that was cooling as a result of the expansion of space \citep[e.g.,][]{haiman1997}, and produced the first black holes \citep[e.g.,][]{madau2001,micic2011}. The final collapse of single rapidly rotating massive stars in low-metallicity environments is a potential channel toward the production of hypernovae and long-duration gamma-ray bursts \citep[e.g.,][]{yoon2005, woosley2006}. \par The study of low-metallicity massive stars is thus crucial in our understanding of the early Universe. While the Magellanic Clouds provide access to massive stars in environments with metallicities down to 20\% of solar, for lower metallicities we have to look to more pristine dwarf galaxies in the Local Group. With 8-10m class telescopes, the stellar populations in these galaxies can be resolved, but obtaining spectra of individual massive stars hosted by these systems remains challenging and expensive in terms of observing time. Consequently, this has so far mostly been done at low spectral resolution \citep[at resolving power $R = \lambda / \Delta \lambda \sim 1\,000-2\,000$; e.g., ][]{bresolin2006, bresolin2007, evans2007, castro2008}. \par The advent of X-Shooter \citep{vernet2011} on ESO's Very Large Telescope (VLT) has opened up the opportunity to observe massive stars in galaxies as far as the edge of the Local Group at intermediate resolution \citep[$R \sim 5\,000-11\,000$,][]{hartoog2012}. Apart from the better resolved shapes of the spectral lines, a higher spectral resolution facilitates a better nebular subtraction. This allows for a more detailed quantitative spectroscopic analysis. \par As the mass loss of massive stars through their stellar winds dominates their evolution, understanding the physical mechanism driving these winds is very important. The winds are thought to be driven by radiation pressure on metallic ion lines \citep[e.g.,][]{lucy1970, castor1975,kudritzki2000}. Consequently, the strength of the stellar winds is expected to scale with metallicity, with the prediction that $\dot{M} \propto Z^{0.69\pm0.10}$ \citep{vink2001}. This metallicity scaling has been verified empirically by \cite{mokiem2007}, who find $\dot{M} \propto Z^{0.78\pm0.17}$ for O stars in the Galaxy and Magellanic Clouds. \begin{table} \centering \caption{Adopted properties of the host galaxies.}\label{tab:galaxies} \begin{tabular}{l c c c c} \hline\hline Galaxy & d & $E(B-V)$ & $Z/Z_{\odot}$\tablefootmark{a} & References\\ & {\tiny (kpc)} & \\ \hline \\[-8pt] IC~1613 & 720 & 0.025 & 0.16 & 1, 2, 3\\ WLM & 995 & 0.08 & 0.13 & 4\\ NGC~3109 & 1300 & 0.14 & 0.12 & 5, 6, 7\\ \hline \end{tabular} \tablefoot{ \tablefoottext{a}{Metallicity for IC~1613 and NGC~3109 are derived from B-supergiants and based on the oxygen abundance, adopting $12 + \log(\mathrm{O}/\mathrm{H})_{\odot} = 8.69$ \citep{asplund2009}. WLM metallicity is based on abundances of iron-group elements obtained from B-supergiants.}} \tablebib{ (1) \citet{pietrzynski2006}; (2) \citet{schlegel1998}; (3) \citet{bresolin2007}; (4) \citet{urbaneja2008}; (5) \citet{soszynski2006}; (6) \citet{davidge1993}; (7) \citet{evans2007}} \end{table} \begin{figure}[!h] \resizebox{\hsize}{!}{\includegraphics{IC1613.pdf}} \caption{Location of the target stars in IC~1613. North is up and east to the left.} \label{fig:ic1613} \resizebox{\hsize}{!}{\includegraphics{WLM.pdf}} \caption{Location of the target star in WLM. North is up and east to the left.} \label{fig:wlm} \resizebox{\hsize}{!}{\includegraphics{NGC3109.pdf}} \caption{Location of the target stars in NGC~3109. North is up and east to the left.} \label{fig:ngc3109} \end{figure} \par To quantify the $\dot{M}(Z)$ relation in even lower metallicity environments, we presented the first intermediate-resolution quantitative spectroscopic analysis of O-type stars with a oxygen abundance that suggests a sub-SMC metallicity in \citet[][henceforth Paper I]{tramper2011}. We unexpectedly found stellar winds that are surprisingly strong, reminiscent of an LMC metallicity. This apparent discrepancy with radiation-driven wind theory is strongest for two stars, one in WLM and one in NGC~3109. \cite{herrero2012} also report a stronger than predicted wind strength for an O-type star in IC~1613. However, observations of a larger sample of stars, as well as observations in the UV, are necessary to firmly constrain the wind properties of these stars and to prove or disprove that O stars at low metallicities have stronger winds than anticipated. \par A first step towards this goal has been made by \cite{garcia2014}, who obtained HST-COS spectra of several O-type stars in IC~1613, and used these to derive terminal wind velocities. They show that the wind momentum for the star analysed by \cite{herrero2012} can be reconciled with the theoretical predictions when their empirical value for the terminal velocity is adopted. They also find indications that the $\alpha$-to-iron ratio in IC~1613 may be sub-solar, which could partly explain the observed strong winds. A full analysis of the UV spectrum to constrain the mass-loss properties of the stars in their sample is still to be done. \par In this paper, we extend our optical sample of O stars in low-metallicity galaxies by four. We constrain the physical properties of the full sample of ten O stars and reassess their winds strengths. Furthermore, we discuss the evolutionary state of the objects, that are among the visually brightest of their host galaxies. We use our results in combination with results from the literature to reassess the low-metallicity effective temperature - spectral type scale. \par The location of all stars in our sample within their host galaxies is indicated in Figures~\ref{fig:ic1613}, \ref{fig:wlm} and \ref{fig:ngc3109}. The host galaxies are of a late type (dwarf irregulars), and have likely been forming stars continuously during their life \citep{tolstoy2009}. The distance and metallicity of the host galaxies that we adopt are given in Table~\ref{tab:galaxies} (but see Section~\ref{sec:mdot} for a discussion on the metallicities). \begin{table*} \centering \caption{Observational properties of the target stars.}\label{tab:stars} \begin{tabular}{l l c c c c c c} \hline\hline ID\tablefootmark{a} & ID & R.A. & Decl. & $V$\tablefootmark{b} & Spectral Type & $M_V$ & RV\\ {\tiny \textit{This work}} & {\tiny \textit{Previous}} &(J2000)& (J2000) & & & & km s$^{-1}$\\ \hline \\[-8pt] IC1613-1 (I1)$^{1,3}$ & A13 & 01 05 06.21& +02 10 44.8 & 19.02 & O3.5 V((f)) & $-5.55$ & $-240$\\ IC1613-2 (I2)$^{1,3}$ & A15 & 01 05 08.74 & +02 10 01.1 & 19.35 & O9.5 III & $-5.11$ & $-240$\\ IC1613-3 (I3)$^{1,3}$ & B11 & 01 04 43.82 & +02 06 46.1 & 18.68 & O9.5 I & $-5.84$ & $-240$\\ IC1613-4 (I4)$^{1,3}$ & C9 & 01 04 38.63 & +02 09 44.4 & 19.02 & O8 III((f)) & $-5.44$ & $-265$\\ IC1613-5 (I5)$^{2,3}$ & B7 & 01 05 01.95 & +02 08 06.5 & 18.99 & O9 I & $-5.29$ & $-214$\\ WLM-1 (W1)$^{1,4}$ & A11 & 00 01 59.97 & $-$15 28 19.2 & 18.40 & O9.7 Ia & $-6.83$\tablefootmark{c} & $-135$\\ NGC3109-1 (N1)$^{1,5}$ & 20 & 10 03 03.22 & $-$26 09 21.4 & 19.33 & O8 I & $-6.67$ & $407$\\ NGC3109-2 (N2)$^{2,5}$ & 33 & 10 03 02.45 & $-$26 09 36.11 & 19.57& O9 If & $-6.41$ & $504$ \\ NGC3109-3 (N3)$^{2,5}$ & 34 & 10 03 14.24 & $-$26 09 16.96 & 19.61 & O8 I(f) & $-6.39$ & $415$\\ NGC3109-4 (N4)$^{2,5}$ & 35 & 10 03 13.65 & $-$26 09 55.76 & 19.70 & O8 I(f) & $-6.28$ & $386$\\ \hline \end{tabular} \tablefoot{ \tablefoottext{a}{In some figures we use the short notation between brackets.} \tablefoottext{b}{$V$-magnitudes from (3),(4), and (5).} \tablefoottext{c}{The value of $-6.35$ listed in Paper I is a typo. The correct value was used in the analysis.}} \tablebib{ (1) Paper I; (2) This work; (3) \citet{bresolin2007}; (4) \citet{bresolin2006}; (5) \citet{evans2007}. } \end{table*} \par In the next section we give an overview of the observations and the data reduction. In Section~\ref{sec:analysis} we describe the analysis and present the results. We discuss the low-metallicity effective temperature scale in Section~\ref{sec:teff}, and the wind strengths in Section~\ref{sec:mdot}. Finally, we discuss the evolutionary properties of the sample and the recent star formation history of the host galaxies in Section~\ref{sec:discussionproperties}. We summarize our findings in Section~\ref{sec:summary}. \section{Observations and data reduction}\label{sec:obs} All stars have been observed with X-Shooter \citep{vernet2011} at ESO's Very Large Telescope as part of the NOVA program for guaranteed time observations. An overview of the observational properties of the stars is given in Table~\ref{tab:stars}. Throughout this paper, we will use the identification given in this table. \par The observations and data reduction of IC1613-1 to 3, WLM-1 and NGC3109-1 (program ID 085.D-0741) are described in Paper~I. An overview of observations of the other stars (under program IDs 088.D-0181 and 090.D-0212) is given in Table~\ref{tab:observations}. All stars were observed with a slit width of 0.8", 0.9" and 0.9" in the UVB, VIS and NIR arms, respectively. The corresponding resolving power $R$ is 6\,200 (UVB), 7\,450 (VIS) and 5\,300 (NIR). All observations were carried out while the moon was below the horizon or illuminated less than 30\% (dark conditions). The data reduction of the newly observed stars was performed with the X-Shooter pipeline v2.2.0. To obtain uncontaminated 1D spectra, the science reduction was done without sky subtraction for each individual exposure. The resulting 2D spectra were folded in the wavelength direction and inspected for the presence of other objects in the slit. A clean part of the slit was then used for sky subtraction. The 1D spectra were extracted from the sky-subtracted 2D spectra. As the observed spectra suffer from nebular emission in the hydrogen and \ion{He}{I} lines, the extracted spectra were carefully inspected for residuals of nebular lines. If needed, a more suitable part of the slit was used. Whenever residuals remained after this procedure, they were clipped from the spectrum before the analysis. \par The 1D spectra of the individual exposures were combined by taking the median flux at each wavelength so cosmic ray hits are removed. Finally, the extracted 1D spectra were normalized by fitting a 4th degree polynomial to the continuum, and dividing the flux by this function. Figure~\ref{fig:atlas} shows the resulting normalized spectra of all stars. \par The spectra have a signal-to-noise ratio (S/N) between 25 and 45 per wavelength bin of 0.2 \AA \ in the UVB\footnote{Note that we are oversampling the spectral resolution. A S/N of 25 per wavelength bin of 0.2 \AA \ corresponds to a S/N of $\simeq$ 50 per resolution element at 4500 \AA. }. As expected in O stars, all the spectra show strong hydrogen, \ion{He}{i} and \ion{He}{ii} lines. Some spectra also show weak nitrogen lines. \section{Analysis \& results}\label{sec:analysis} To investigate the properties of the target stars, we first obtained the stellar and wind parameters by fitting synthetic spectra to the observed line profiles. The method is described in the following section, and the results are presented in Section~\ref{sec:GAresults}. The stellar parameters were then used to obtain estimates of the evolutionary parameters (Section~\ref{sec:bonnsai}). We comment on the results of the individual targets in Section\ref{sec:individual}. \begin{table} \centering \caption{Journal of observations.}\label{tab:observations} \begin{tabular}{l l c c c} \hline\hline ID & HJD & $t_{\mathrm{exp}}$ & \multicolumn{2}{c}{Average seeing} \\ & {\tiny \textit{At start of obs.}}& (s) & UVB (")& VIS (") \\ \hline \\[-8pt] IC1613-5 & 2\,455\,858.653 & 4x900 & 1.1 & 1.0 \\ & 2\,455\,858.705 & 2x900 & 1.1 & 0.9 \\ NGC3109-2 & 2\,456\,337.531 & 4x900 & 2.2 & 1.2\\ & 2\,456\,337.587 & 4x900 & 1.4 & 1.0\\ & 2\,456\,338.583 & 4x900 & 1.4& 1.0\\ & 2\,456\,338.638 & 4x1200 & 1.0 & 0.8\\ NGC3109-3 & 2\,456\,337.643 & 4x1100 & 1.1& 0.8\\ & 2\,456\,337.732 & 4x1100 & 0.9 & 0.6\\ & 2\,456\,338.736 & 4x900 & 0.8 & 0.6\\ & 2\,456\,338.791 & 2x900 & 1.0 & 0.6\\ NGC3109-4 & 2\,456\,337.797 & 4x900 & 1.2& 0.7\\ & 2\,456\,337.853 & 2x1200 & 1.7 & 0.7\\ & 2\,456\,338.823 & 4x1200 & 1.5 & 0.7\\ \hline \end{tabular} \end{table} \subsection{Fitting method}\label{sec:GA} \begin{figure*} \resizebox{\hsize}{!}{\includegraphics{atlas.pdf}} \caption{Observed spectra (black dots) and best-fit line profiles (red lines). Rest wavelengths of the fitted spectral lines are indicated by the vertical dashed lines. In this plot the wavelength has been corrected for the radial velocities listed in Table~\ref{tab:stars}, and binned to 0.5 \AA.} \label{fig:atlas} \end{figure*} To determine the stellar and wind properties, we used an automated fitting method developed by \cite{mokiem2005}. This method fits spectra produced by the non-LTE model atmosphere code {\sc fastwind} \citep{puls2005} to the observed spectrum using the genetic algorithm based fitting routine {\sc pikaia} \citep{charbonneau1995}. This genetic algorithm (GA) method allows for a thorough exploration of parameter space in affordable CPU time on a supercomputer. \par The absolute $V$-band magnitude $(M_\mathrm{V})$ is needed as input for the GA in order to determine the luminosity (Table~\ref{tab:stars}). $M_V$ was calculated using the $V$ magnitudes also given in Table~\ref{tab:stars} and distances and mean reddening listed in Table~\ref{tab:galaxies}. \par The radial velocity (RV) of each star is also listed in Table~\ref{tab:stars}. These were measured by fitting Gaussians to the \ion{H}{$\gamma$} and \ion{He}{i} $\lambda$4471 lines and calculating the average velocity needed to match the observed wavelength shifts. \par The parameters that are obtained from the atmosphere fitting are the effective temperature $(T_{\mathrm{eff}})$, the surface gravity $(g)$, the mass-loss rate $(\dot{M})$, the surface helium abundance $(N_{\mathrm{He}})$, the atmospheric microturbulent velocity $(v_{\mathrm{tur}})$ and the projected rotational velocity $(v_{\mathrm{rot}}\sin{i})$. As in Paper I, the parameter describing the rate of acceleration of the outflow $(\beta)$ can not be constrained from the data, and was fixed to the value predicted by theory \citep[$\beta=0.95$ for the supergiants presented in this work;][]{muijres2012}. \par The terminal wind velocity $(v_{\infty})$ can not be constrained from the optical spectrum. Therefore, we used the empirical scaling with the escape velocity ($v_{\mathrm{esc}}$) for Galactic stars \citep[$v_{\infty}=2.65\,v_{\mathrm{esc}}$;][]{kudritzki2000}, and scaled these using \citet[][$v_{\infty} \propto Z^{0.13}$]{leitherer1992} to correct for the lower metallicity. In contrast to Paper I, this metallicity scaling has now been implemented into the GA, and was applied before running each individual {\sc fastwind} model. Therefore, the mass-loss rate no longer has to be scaled down after the fitting, as was previously needed. \par Several additional small changes were implemented in the GA: \begin{itemize} \item{The best-fitting model is now selected based on the $\chi^2$, which is also used for the error calculation.} \item{\ion{He}{i} $\lambda$4922 is now also fitted in addition to the 11 lines that were used in Paper I.} \item{The minimum microturbulent velocity $v_{\mathrm{tur}}$ is set to 5 instead of 0, as {\sc fastwind} models with $v_{\mathrm{tur}} < 5$ may not be accurate.} \item{The error on the flux is now based on the signal-to-noise ratio (S/N) calculated near each of the fitted lines, instead of a single value for each X-Shooter arm.} \end{itemize} To present a homogeneous analysis, we have re-analysed the stars of Paper I with the updated fitting routine, and we present the new parameters here. In general, the new values are in excellent agreement with those presented in Paper I. The exception is IC1613-A1, where \ion{H}{$\alpha$} was not properly normalized in Paper I. For this star we re-normalized \ion{H}{$\alpha$}, resulting in a somewhat higher mass-loss rate. \par Table~\ref{tab:bestfit} presents the best-fit parameters for each of the target stars. The synthetic spectra of the corresponding {\sc fastwind} models are overplotted on the observed spectra in Figure~\ref{fig:atlas}. \subsection{Derived properties and error calculation}\label{sec:GAresults} \begin{table*} \centering \caption{Best-fitting stellar and wind parameters.}\label{tab:bestfit} \begin{tabular}{l c c c c c c} \hline\hline ID & $T_{\mathrm{eff}}$ & $\log{g}$ & $\log{\dot{M}}$ & $N_{\mathrm{He}}/N_{\mathrm{H}}$ & $v_{\mathrm{tur}}$ & $v_\mathrm{rot} \sin{i}$ \\ & \tiny{(kK)} & \tiny{(cm s$^{-2}$)} & \tiny{($M_{\odot}$ yr$^{-1}$)} & & \tiny{(km s$^{-1}$)} &\tiny{(km s$^{-1}$)} \\ \hline \\[-8pt] IC1613-1 & $45.40^{+2.00}_{-2.25}$ & $3.65^{+0.16}_{-0.10}$ & $-5.85^{+0.10}_{-0.50}$ & $0.25^{\uparrow}_{0.10}$ & $24^{+6}_{\downarrow}$ & $98^{+36}_{-44}$ \\ IC1613-2 & $33.85^{+2.10}_{-2.75}$ & $3.77^{+0.33}_{-0.33}$ & $-6.35^{+0.35}_{\downarrow}$ & $0.14^{+0.15}_{-0.08}$ & $11^{+15}_{\downarrow}$ & $32^{+38}_{-22}$ \\ IC1613-3 & $31.45^{+1.65}_{-2.45}$ & $3.41^{+0.22}_{-0.19}$ & $-6.25^{+0.35}_{-1.20}$ & $0.15^{+0.09}_{-0.08}$ & $5^{+18}_{\downarrow}$ & $94^{+32}_{-24}$ \\ IC1613-4 & $35.2^{+1.85}_{-1.40}$ & $3.52^{+0.20}_{-0.11}$ & $-6.25^{+0.15}_{-0.50}$ & $0.12^{+0.10}_{-0.04}$ & $17^{+7}_{\downarrow}$ & $76^{+16}_{-26}$ \\ IC1613-5 & $35.05^{+4.55}_{-4.8}$ & $3.74^{\uparrow}_{-0.44}$ & $-$ & $0.06^{+0.17}_{\downarrow}$ & $27^{\uparrow}_{\downarrow}$ & $270^{+112}_{-92}$\\ WLM-1 & $30.60^{+1.70}_{-3.60}$ & $3.28^{+0.19}_{-0.31}$ & $-5.50^{+0.10}_{-0.35}$ & $0.22^{\uparrow}_{0.10}$ & $9^{+9}_{\downarrow}$ & $72^{+36}_{-22}$ \\ NGC3109-1 & $35.15^{+3.20}_{-2.55}$ & $3.53^{+0.30}_{-0.42}$ & $-5.35^{+0.15}_{-0.35}$ & $0.09^{0.22}_{0.04}$ & $17^{\uparrow}_{\downarrow}$ & $110^{+50}_{-52}$ \\ NGC3109-2 & $33.30^{+3.30}_{-2.25}$ & $3.35^{+0.43}_{-0.19}$ & $-5.45^{+0.30}_{-0.15}$ & $0.09^{+0.18}_{-0.05}$ & $10^{\uparrow}_{\downarrow}$ & $200^{+102}_{-104}$ \\ NGC3109-3 & $33.05^{+1.45}_{-1.25}$ & $3.16^{+0.19}_{-0.11}$ & $-5.75^{+0.15}_{-0.25}$ & $0.12^{+0.17}_{-0.03}$ & $23^{+3}_{\downarrow}$ & $82^{+30}_{-32}$\\ NGC3109-4 & $35.05^{+3.45}_{-4.45}$ & $3.43^{+0.51}_{-0.31}$ & $-5.55^{+0.25}_{-0.45}$ & $0.08^{+0.15}_{-0.04}$ & $25^{+5}_{\downarrow}$ & $96^{+60}_{-70}$\\ \hline \end{tabular} \end{table*} \begin{table*} \centering \caption{Properties derived from best-fit parameters.}\label{tab:derivedproperties} \begin{tabular}{l c c c c c} \hline\hline ID &$v _{\infty}$ & $\log{L}$ & $R$ & $M_{\mathrm{spec}}$ & $\log{D_\mathrm{mom}}$ \\ & \tiny{(km\,s$^{-1}$)} & \tiny{($L_{\odot}$)} & \tiny{($R_{\odot}$)} & \tiny{($M_{\odot}$)} & \tiny{(g\,cm\,s$^{-2}$\,$R_{\odot}^{1/2}$)}\\ \hline \\[-8pt] IC1613-1 & $1755^{+328}_{-165}$ & $5.71^{+0.05}_{-0.06}$ & $11.9^{+0.4}_{-0.3}$ & $22.6^{+8.4}_{-3.6}$ & $28.73^{+0.11}_{-0.53}$ \\ IC1613-2 & $2022^{+872}_{-593}$ & $5.21^{+0.06}_{-0.09}$ & $11.9^{+0.8}_{-0.5}$ & $30.2^{+29.0}_{-14.5}$ & $28.29^{+0.46}_{\downarrow}$ \\ IC1613-3 & $1625^{+440}_{-296}$ & $5.42^{+0.06}_{-0.08}$ & $17.6^{+1.2}_{-0.7}$ & $28.9^{+16.4}_{-8.9}$ & $28.38^{+0.42}_{-1.19}$ \\ IC1613-4 & $1558^{+387}_{-184}$ & $5.32^{+0.06}_{-0.04}$ & $12.5^{+0.4}_{-0.5}$ & $18.9^{+10.1}_{-4.2}$ & $28.29^{+0.19}_{-0.51}$ \\ IC1613-5 & $2010^{+2672}_{-742}$ & $5.32^{+0.13}_{-0.16}$ & $12.6^{+1.5}_{-1.1}$ & $31.6^{+129.8}_{-18.4}$ & $-$\\ WLM-1 & $1777^{+435}_{-488}$ & $5.79^{+0.06}_{-0.13}$ & $28.3^{+3.0}_{-1.2}$ & $55.6^{+30.5}_{-24.3}$ & $29.28^{+0.14}_{-0.45}$ \\ NGC3109-1 & $2166^{+876}_{-1374}$ & $5.87^{+0.10}_{-0.08}$ & $23.7^{+1.4}_{-1.5}$ & $69.1^{+65.3}_{-39.7}$ & $29.47^{+0.28}_{-0.53}$ \\ NGC3109-2 & $1692^{+1058}_{-331}$ & $5.71^{+0.10}_{-0.13}$ & $21.9^{+1.2}_{-1.5}$ & $38.9^{+65.9}_{-13.7}$ & $29.25^{+0.51}_{-0.24}$\\ NGC3109-3 & $1357^{+328}_{-156}$& $5.69^{+0.05}_{-0.04}$ & $21.8^{+0.7}_{-0.7}$ & $25.0^{+13.3}_{-5.32}$ & $28.85^{+0.24}_{-0.26}$\\ NGC3109-4 & $1767^{+1412}_{-492}$ & $5.71^{+0.11}_{-0.14}$ & $19.8^{+0.5}_{-1.3}$ & $38.5^{+86.0}_{-17.2}$ & $29.15^{+0.45}_{-0.56}$\\ \hline \end{tabular} \end{table*} \par In addition to $v_{\infty}$ (discussed above), several important quantities can be derived from the best-fit parameters: the bolometric luminosity ($L$), the stellar radius ($R$), and the spectroscopic mass ($M_{\mathrm{spec}}$). These are given in Table~\ref{tab:derivedproperties}. This table also gives the modified wind momentum, which is defined as $D_{\mathrm{mom}} = \dot{M} \, v_{\infty} \, \sqrt{R/R_{\odot}}$ and is ideal to study the mass loss as it is almost independent of mass. Furthermore, $D_{\mathrm{mom}}$ scales with the luminosity through $R$, making it less sensitive to uncertainties in the luminosity determination. \par To derive the error bars given in Tables~\ref{tab:bestfit} and \ref{tab:derivedproperties}, we first divide all $\chi^2$ values with a factor such that the best model has a reduced $\chi^2$ of unity. This ensures meaningful error bars that are not influenced by under or overestimated errors on the flux (see Paper I). We then calculate the probability $P = 1 - \Gamma(\chi^2/2, \nu/2)$, with $\Gamma$ the incomplete gamma function and $\nu$ the degrees of freedom, for each model. $P$ quantifies the probability that a $\chi^2$ value differs from the best-fit $\chi^2$ due to random fluctuations. Models that satisfy $P \geq 0.05$ are accepted as providing a suitable fit, and the range covered by the stellar parameters of these models (and the properties derived from them) is taken as the 95\% confidence interval. This method also ensures that uncertainties in the parameters that arise due to clipped parts of the spectrum are reflected in the error bars. \subsection{Mass and age}\label{sec:bonnsai} To determine the evolutionary parameters of the stars, we use the {\sc bonnsai}\footnote{The {\sc bonnsai} web-service is available at bonnsai.astro.uni-bonn.de.} tool (Schneider et al. in prep.). {\sc bonnsai} uses Bayes' theorem to constrain key stellar parameters, such as initial mass and age, by comparing the observed stellar parameters to theoretical predictions from stellar evolution. \par We obtained an estimate of the initial mass $(M_{\mathrm{ini}})$, the current mass $(M_{\mathrm{act}})$, the initial and current rotation ($v_{\mathrm{rot,ini}}$ and $v_{\mathrm{rot,act}}$), and the age of the stars. We use the evolutionary tracks for SMC metallicity of \cite{brott2011}, as these are closest in metallicity. As we do not find a significant difference in the temperature of our stars compared to similar SMC stars (see Section~\ref{sec:teff}), the use of the SMC tracks does not induce large systemetic uncertainties in the evolutionary parameters. As priors to the {\sc bonnsai} method we choose a \cite{salpeter1955} initial mass function, and the \cite{ramirez2013} 30 Doradus distribution for the initial rotational velocity. \par As input observables we used the luminosity, effective temperature, surface gravity and projected rotational velocity. {\sc bonnsai} adapts these parameters based on the comparison with the evolutionary predictions. The posterior reproduced parameters are within errors of the input values. The estimated evolutionary parameters are given in Table~\ref{tab:bonnsai}. The stellar masses that are derived with the {\sc bonnsai} method are in good agreement with the mass estimates that would be derived using the conventional method, i.e. a visual comparison with evolutionary tracks in the Herzsprung-Russell diagram (HRD, Figure~\ref{fig:hrd}). \begin{table*} \centering \caption{Stellar parameters obtained from comparison with evolutionary tracks using {\sc Bonnsai}}.\label{tab:bonnsai} \begin{tabular}{l c c c c c} \hline\hline ID & $M_{\mathrm{ini}}$ & $M_{\mathrm{act}}$ & $v_{\mathrm{rot,ini}}$ & $v_{\mathrm{rot, act}}$ & $\tau$ \\ & \tiny{($M_{\odot}$)} & \tiny{($M_{\odot}$)} & \tiny{(km\,s$^{-1}$)} & \tiny{(km\,s$^{-1}$)} & \tiny{(Myr)} \\ \hline \\[-8pt] IC1613-1 & $49.0^{+3.5}_{-3.4}$ & $47.6^{+3.6}_{-3.1}$ & $100^{+48}_{-40}$ & $100^{+80}_{-54}$ & $2.32^{+0.31}_{-0.34}$ \\ IC1613-2 & $24.6^{+2.1}_{-1.9}$ & $24.4^{+1.9}_{-1.8}$ & $70^{+41}_{-38}$ & $70^{+41}_{-38}$ & $5.00^{+0.71}_{-0.62}$ \\ IC1613-3 & $29.4^{+2.4}_{-2.3}$ & $28.8^{+2.2}_{-2.2}$ & $110^{+40}_{-36}$ & $100^{+49}_{-27}$ & $4.74^{+0.45}_{-0.36}$ \\ IC1613-4 & $28.8^{+1.8}_{-1.4}$ & $28.4^{+1.7}_{-1.3}$ & $80^{+37}_{-28}$ & $80^{+37}_{-28}$ & $4.40^{+0.32}_{-0.36}$ \\ WLM-1 & $41.6^{+6.4}_{-5.3}$ & $39.8^{+6.1}_{-4.7}$ & $90^{+48}_{-29}$ & $90^{+45}_{-32}$ & $3.58^{+0.49}_{-0.33}$ \\ NGC3109-1 & $52.6^{+5.1}_{-4.3}$ & $50.0^{+5.1}_{-3.6}$ & $110^{+59}_{-45}$ & $110^{+59}_{-46}$ & $2.86^{+0.29}_{-0.22}$ \\ NGC3109-2 & $40.0^{+6.4}_{-5.4}$ & $38.6^{+6.0}_{-5.0}$ & $130^{+111}_{-61}$ & $130^{+113}_{-63}$ & $3.44^{+0.50}_{-0.41}$ \\ NGC3109-3 & $42.0^{+2.7}_{-2.3}$ & $40.4^{+2.6}_{-2.0}$ & $90^{+48}_{-31}$ & $90^{+48}_{-31}$ & $3.54^{+0.19}_{-0.17}$ \\ NGC3109-4 & $39.8^{+6.9}_{-5.9}$ & $38.6^{+6.5}_{-5.4}$ & $100^{+62}_{-47}$ & $100^{+63}_{-48}$ & $3.34^{+0.58}_{-0.47}$ \\ \hline \end{tabular} \end{table*} \subsection{Comments on individual stars}\label{sec:individual} \begin{table*} \centering \caption{Coefficients for the spectral type - $T_{\mathrm{eff}}$ calibrations.}\label{tab:SpT_Teff} \begin{tabular}{l c c c c} \hline\hline & \multicolumn{2}{c}{Unweighted} & \multicolumn{2}{c}{Weighted} \\ Sample & a & b & a & b \\ \hline\\[-8pt] This work & 47584.3 & $-$1595.4 & 44907.9$\pm$4457.2 & $-$1339.1$\pm$518.9 \\ Low Z & 53398.0 & $-$2203.6 & 56636.1$\pm$766.9 & $-$2677.4$\pm$106.1 \\ Low Z (no O3) & 48670.7 & $-$1653.2 & 47283.6$\pm$3017.7 & $-$1588.2$\pm$356.1 \\ SMC & 51929.7 & $-$2138.8 & 50189.7$\pm$1329.2 & $-$1957.2$\pm$167.4 \\ \hline \end{tabular} \tablefoot{Here, $a$ and $b$ are the coefficients in $T_{\mathrm{eff}} = a + b \times X_{\mathrm{SpT}}$, with $X_{\mathrm{SpT}}$ the O subtype.} \end{table*} \subsubsection{IC1613-1} This is the only dwarf star in the sample, which is reflected by its young derived age (Table~\ref{tab:bonnsai}). In Paper I, this was the only star with a wind momentum lower than the empirical SMC values from \cite{mokiem2007}. However, as already mentioned, \ion{H}{$\alpha$} was not properly normalized, which caused the mass-loss rate to be slightly underestimated. In our new analysis of this star we renormalized \ion{H}{$\alpha$}, and the updated modified wind momentum is now comparable to those found for SMC stars. \par \cite{garcia2014} obtained the UV spectrum of IC1613-1 using HST-COS, and used it to determine the terminal wind velocity. They find $v_{\infty} = 2\,200^{+150}_{-100}$ km s$^{-1}$, somewhat higher than the values of 1\,869 km s$^{-1}$ (Paper I) and 1\,755 km s$^{-1}$ (this work) that we obtain from the scaling with the escape velocity. Their value for the terminal wind velocity would result in a value of $\log(D_{\mathrm{mom}})$ that is $\simeq 0.1$ dex higher than ours. \par IC1613-1 has a low surface gravity for its luminosity class, and is enriched in helium. \subsubsection{IC1613-2} The nebular emission is variable along the X-shooter slit, which prevents a good nebular subtraction. As a consequence, a large part of the core of the Balmer lines had to be clipped from the spectrum before fitting. Without the core of \ion{H}{$\alpha$} we can only derive an upper limit for the mass-loss rate of this star. \subsubsection{IC1613-3 and IC1613-4} Both these stars are well fitted by the atmosphere models. For both stars \ion{H}{$\alpha$} is strongly in absorption and the mass-loss rate cannot be well constrained from this line. This results in fairly large error bars on their modified wind momenta. \subsubsection{IC1613-5} After our observations of this object, it was found to be an eclipsing binary \citep{bonanos2013}. Our spectra show strong variability in \ion{He}{ii} $\lambda$4686 and \ion{H}{$\alpha$} between individual exposures, which may be due to colliding winds \citep{stevens1992}. Although we provide parameters from fitting the other lines, it is likely that the spectrum is composite (depending on the mass ratio). The listed values are therefore only representative of the composite spectrum. This is a possible cause of the broad spectral lines of IC1613-5 (see Figure~\ref{fig:atlas}), although the rotational velocity of $v_{\mathrm{rot}}\sin{i} = 270$ km s$^{-1}$ that is needed to fit these lines is not unphysically high \citep[see][]{ramirez2013}. The variability does prevent us to constrain the mass-loss rate and consequently modified wind momentum of this star. We excluded this star from both the {\sc bonnsai} analysis and our discussion of the mass-loss rates (Section~\ref{sec:mdot}). \subsubsection{WLM-1} This is the only star in our sample in WLM, and one of the stars from Paper I which showed a large discrepancy with radiation-driven wind theory. The wind properties derived with the updated GA are very similar to those of Paper I. This star also has a high helium abundance. \subsubsection{NGC3109-1} This star shows a strong stellar wind, and is one of the stars in Paper I that exhibits the largest discrepancy with radiation-driven wind theory. While \ion{H}{$\alpha$} is still slightly in absorption, \ion{He}{ii} $\lambda$4686 is fully filled in. \subsubsection{NGC3109-2 and NGC3109-4} Both stars show signs of strong winds in their spectrum. In the fitting, the line center of \ion{H}{$\alpha$} in NGC3109-4 was clipped due to nebular contribution that could not be fully corrected for. \ion{He}{ii} $\lambda$4686 is fully filled in in both stars. \ion{H}{$\alpha$} is fully filled in for NGC3109-2 and almost fully filled in for NGC3109-3. Unsurprisingly, the derived wind momenta for both stars are high. \subsubsection{NGC3109-3} \ion{He}{ii} $\lambda$4686 is almost fully filled in, while \ion{H}{$\alpha$} is still mildly in absorption. This results in a modest wind strength. \section{Effective temperature scale}\label{sec:teff} \begin{figure}[!t] \resizebox{\hsize}{!}{\includegraphics{SpT_Teff.pdf}} \caption{Spectral type versus effective temperature calibration for giants and supergiants in low-metallicity environments. Symbol size indicates the luminosity class, with the larger symbols for supergiants. Plotted are results from this work, the \cite{herrero2012} and \cite{garcia2013} results for IC1613, and the \cite{mokiem2006} and \cite{massey2004, massey2009} results for the SMC. The solid black line indicates a linear fit to the stars from this paper, not including the error bars on $T_{\mathrm{eff}}$. The red solid line is the unweighted linear fit to the SMC stars, and the blue solid line the fit to the stars in IC~1613, WLM and NGC~3109 (low-Z). The dashed blue line is an unweighted fit to all low-Z results but excluding the single O3 giant. It illustrates the sensitivity of the found relation to this point.} \label{fig:teffcal} \end{figure} \cite{garcia2013} presented the first effective temperature calibration for potentially sub-SMC metallicities (their figure~7). In Figure~\ref{fig:teffcal} we use our results (Table~\ref{tab:bestfit}) to provide an updated version of this calibration. Similar to \cite{garcia2013}, we first determined the spectral type - $T_{\mathrm{eff}}$ relation using an unweighted least-squares linear fit to the temperatures of the giants and supergiants from our work. We did the same for the total sample of low-Z giants and supergiants \citep[this work;][]{herrero2012, garcia2013} and an SMC sample from \cite{mokiem2006} and \cite{massey2004, massey2009}. The coefficients of the derived linear relations are given in Table~\ref{tab:SpT_Teff}. \par The updated $T_{\mathrm{eff}}$ scale for low metalicities is very similar to the relation found by \cite{garcia2013}, and is $\simeq 1000$ K hotter than the SMC relation. This is expected for a lower metallicity, as the stars are hotter due to a slightly smaller radius \citep[the result of lower opacities in the stellar interior; see e.g., ][]{mokiem2004}. However, as \cite{garcia2013} conclude, the significance of the observed difference between the temperature scales is unclear, given the error bars on the temperatures. Also, the low-Z relation is very sensitive to the position of the only O3 III star in IC~1613, in the region of parameter space where the SMC relation is not constrained. This sensitivity is illustrated by the dotted blue line in Figure~\ref{fig:teffcal}, which is the relation found when the O3 star is excluded from the fit. \begin{figure}[!t] \resizebox{\hsize}{!}{\includegraphics{SpT_Teff_errors.pdf}} \caption{Same as Figure~\ref{fig:teffcal}, but with the fitting done including the error bars on $T_{\mathrm{eff}}$. For stars that do not have published error bars, an error of 1 kK was adopted. The shaded areas indicate the uncertainties of the relations found for the SMC and low-metallicity relations.} \label{fig:teffcal_errors} \end{figure} \par As a second step, we included the error bars in our analysis. Figure~\ref{fig:teffcal_errors} presents the relations that are obtained by weighted least-squares linear fits (i.e. including the error bars on $T_{\mathrm{eff}}$) to the same samples. As the error bars on the temperature presented in Table~\ref{tab:bestfit} correspond to the 95\% confidence interval, we use half these values (roughly corresponding to $\sim1 \sigma$ for normally distributed errors). Because symmetric error bars are easier to handle in a simple approach, we use the average of the upper and lower errors. \cite{massey2004, massey2009} do not provide error bars, and we adopt $\pm$1\,000 K for these stars. The coefficients of the relations that we obtain are given in Table~\ref{tab:SpT_Teff}. \par The low-metallicity temperature scale obtained from the weighted fits is steeper than the SMC scale, and no longer above it at each point of the spectral range. The error bars on both relations overlap over the entire spectral range covered by the SMC stars. Additionally, the effective temperature at spectral type O8 obtained from our sample, which is well constrained by four stars, is very close to the SMC value regardless of the fitting method. \par Thus, with the number of stars that are currently analysed we do not find a significant difference between the effective temperature calibrations for the host galaxies of the stars studied in this paper and the SMC. However, a good comparison is hampered by the small sample size, and the absence of early-type giants and supergiants in the SMC sample. Ideally, the low-metallicity effective temperature scale has to be derived from a large number of dwarfs of all subtypes, which are also found in the SMC. However, even if a sufficient number of O-type dwarfs is present in the low-metallicity galaxies, obtaining their spectra will have to await the advent of 30m-class telescopes. \section{Mass loss versus metallicity}\label{sec:mdot} \begin{figure}[!t] \resizebox{\hsize}{!}{\includegraphics{wld.pdf}} \caption{Location of the target stars in the modified wind momentum versus luminosity diagram. Also indicated are the theoretical predictions from \cite{vink2001} and the empirical results from \cite{mokiem2007}. NGC3109-4 is shifted by $-$0.01 dex in luminosity for clarity. $1 \sigma$ error bars are indicated. The thick line represents a linear fit to our results, and shows that the wind strengths are comparable to the empirical LMC results.} \label{fig:wld} \end{figure} In Paper I, we reported that the wind momenta of the stars in our sample appear to be higher than theoretically predicted for their metallicity. This trend remains after refitting these stars with the updated GA and including the new targets. This is shown in the updated modified wind momentum versus luminosity diagram (WLD; Figure~\ref{fig:wld}). This figure also shows a weighted linear fit to our data ($\log{D_{\mathrm{mom}}} = a + b \times \log{L/L_{\odot}}$, with $a=18.4 \pm 1.9$, $b=1.86 \pm 0.33$). The fit confirms that the stars exhibit LMC strength winds. Only two stars (IC1613-1 and NGC3109-3) have a best-fit value for their modified wind momentum that is close to SMC values, and none have values indicative of a sub-SMC wind strength. \par An important aspect to note when using the WLD to compare mass-loss rates, is that inhomogeneities in the wind (clumping) are not taken into account when deriving the empirical mass-loss rate. This neglect of clumping causes mass-loss rates derived from diagnostic lines sensitive to the density-squared, such as \ion{He}{ii} $\lambda$4686 and \ion{H}{$\alpha$}, to be over-estimated \citep[e.g.,][]{puls2008}. \par This effect can be seen in Figure~\ref{fig:wld} by comparing the results from \cite{mokiem2007} to the predictions from \cite{vink2001}. The empirical values for the Galaxy, LMC and SMC are clearly higher than the ones predicted by theory. However, the trend of decreasing wind strength at lower metallicities is in excellent agreement with theory. Thus, for an assumed sub-SMC metallicity we would expect our stars to be located below the empirical SMC values in the WLD. \par \cite{lucy2012} argues that the neglect of wind clumping is the most likely explanation for the high mass-loss rates of these stars. However, we do argue that wind clumping would have to be metallicity dependant to explain our results. An other possibility given by \cite{lucy2012} is the presence of an additional wind-driving mechanism, possibly only operating in winds that have low terminal wind velocities, or in a restricted part of $(T_{\mathrm{eff}}, g, Z)$-space. \par However, an explanation for the high mass-loss rates may be found in the assumed metallicity of the host galaxies. Iron (and iron-like elements) remains the dominant element in driving the wind for metallicities down to $0.1 Z_{\odot}$, while $\alpha$ elements dominate at lower metallicities \citep{vink2001}. Thus, the iron content of the stars needs to be evaluated to be able to properly compare the wind strengths with theoretical predictions. \par While all the host galaxies have very low average stellar iron abundances of $[\mathrm{Fe}/\mathrm{H}]\la -1.2$ \citep[for an overview, see][]{mcconnachie2012}, the metallicity of the young stellar population is likely higher. This metallicity can be constrained indirectly from \ion{H}{ii} regions and directly from red and blue supergiants. \cite{garcia2014}, \cite{levesque2012}, and \cite{evans2007} give overviews of all relevant metallicity measurements of the young stellar population of IC~1613, WLM, and NGC3109, respectively. The metalicity measurements range from $0.05\, Z_{\odot}$ to $0.10\, Z_{\odot}$ based on the oxygen abundance in \ion{H}{ii} regions and up to $0.15\, Z_{\odot}$ in blue supergiants. However, there are indications that the galaxies have a sub-solar $\alpha$-to-iron ratio. Below, we give an overview of the {\it stellar} iron (or iron-group elements) abundance measurements in the young population of the host galaxies. \par \cite{tautvaisiene2007} derive an iron content of $[\mathrm{Fe}/\mathrm{H}] = -0.67\pm0.06$ for three M-type supergiants in IC~1613, or $Z_{\mathrm{IC1613}}=0.21\,Z_{\odot}$ based on iron. \cite{garcia2014} find qualitative indications that the iron content might be close to the SMC value. \par \cite{venn2003} report an iron abundance of $[\mathrm{Fe}/\mathrm{H}] = -0.38\pm0.29$ for two supergiants in WLM, corresponding to $Z_{\mathrm{WLM}}=0.42\,Z_{\odot}$, but with very large error bars. They derive a stellar oxygen abundance that is five times higher than those found from nebular studies. Conversely, \cite{urbaneja2008} derive $Z_{\mathrm{WLM}}=0.13\,Z_{\odot}$ based on mainly iron, chromium and titanium in blue supergiants, and find no indication that the $\alpha$-to-iron ratio is non-solar. In particular, they derive a metallicity of $[Z] = -0.80 \pm 0.20$ for WLM-1 ($0.16\,Z_{\odot}$). It therefore seems unlikely that an underestimated iron abundance explains the strong stellar wind of WLM-1. \par \cite{hosek2014} analysed 12 late-B and early-A supergiants in NGC3109, and derive $[Z] = -0.67 +/- 0.13$ based on iron-group elements, or $Z_{\mathrm{NGC3109}}=0.21\pm0.08\,Z_{\odot}$. As for IC~1613, this indicates that the iron content is SMC-like. Our results should thus be compared to the SMC predictions. \par For our sample stars, an SMC metallicity would lessen the discrepancy between the observed wind momenta and those predicted from theory. Compared to the SMC predictions, IC1613-1 is in good agreement with the radiation-driven wind theory, while the other three stars in IC1613 have too high best-fit values but agree within error bars. NGC3109-3 has a slightly too high mass-loss rate but is in agreement within errors. For the other three stars in NGC3109 the best-fit wind strengths are comparable to or slightly higher than LMC values, but can just be reconciled with SMC values within errors for two of them. The wind strength of WLM-1 is just in agreement with an SMC metallicity, but as mentioned above, it is unlikely that the metallicity is underestimated for this star. \par Considering the sample as a whole, the observed discrepancy with radiation-driven wind theory at low metallicities may be reduced if the metallicity has indeed been underestimated. However, the stars still tend to have too high mass-loss rates, even if their iron content is comparable to SMC stars. For our results to be fully in agreement with the predictions from radiation-driven wind theory, the iron content should be LMC-like, or half solar (see Figure~\ref{fig:wld}). Further constraints of the metal content in the host galaxies would be helpful. Most importantly, a confirmation of the wind properties from the UV has to be obtained to reduce the uncertainties in the derived wind momenta. \section{Evolutionary state }\label{sec:discussionproperties} \begin{figure}[!t] \resizebox{\hsize}{!}{\includegraphics{hrd.pdf}} \caption{Herzsprung-Russell diagram indicating the location of the target stars. Also plotted are evolutionary tracks from \cite{brott2011} for SMC metallicity and no initial rotation. The dashed lines indicate isochrones in steps of 1 Myr. The dashed-dotted lines indicate the magnitude cut-offs for the three galaxies, and the dotted vertical line roughly indicates the division between O and B stars. The IC1613 stars from \cite{garcia2013} are plotted in red.} \label{fig:hrd} \end{figure} \par Figure~\ref{fig:hrd} shows the position of the full sample of stars in the Herzsprung-Russell diagram (HRD). Our sample is complete for the O stars listed in \citet{bresolin2006, bresolin2007}, and \citet{evans2007} above the indicated magnitude cut-off for each galaxy. \cite{garcia2013} identified eight new O-type stars in IC1613, and provided an estimate of their stellar parameters. We also show these stars in Figure~\ref{fig:hrd}, but note that they are based on observations with a lower resolving power ($R=1000$). \cite{garcia2013} do not give bolometric luminosities, and the values used in the HRD are based on their temperatures and the bolometric correction from \cite{martins2005}. \par The single WLM star in our sample is at the location in the HRD that is expected for its spectral type. It is remarkable that no other O stars are known in WLM that populate the area of the HRD below WLM-1 and above our magnitude cut-off (indicated with the dashed line in Figure~\ref{fig:hrd}). The only other known O star in WLM is an O7 V((f)) with V=20.36 \citep[A15 in][]{bresolin2006}. This suggests that, while star formation is ongoing in WLM, this mostly happens in low-mass clusters that do not produce many O-type stars. \par For NGC~3109, our sample is restricted by the magnitude cut-off. The stars in our sample populate the small area of the HRD that we can observe with X-Shooter, and thus all have high masses ($M_{\mathrm{ini}} \ga 40 M_{\odot}$). They are located in different regions within the host galaxy (see Figure~\ref{fig:ngc3109}), suggesting that massive star formation is ongoing in several regions of the galaxy. \par The HRD for IC~1613 is well populated by our sample and the stars from \cite{garcia2013}. Most of the stars have masses in the range $25 M_{\odot} \la M_{\mathrm{ini}} \la 35 M_{\odot}$, but the two O3 stars indicate that higher mass stars are also being formed. This is further suggested by the presence of the oxygen sequence Wolf-Rayet star in the galaxy \citep[DR1; see, e.g., ][]{kingsburgh1995, tramper2013}. While on large time-scales the star-formation rate has been constant \citep{skillman2014}, IC~1613 is currently rigorously forming stars, with 164 OB associations identified \citep{garcia2009}. The location of our sample of stars in IC~1613 follows the main regions of star formation, with the most massive star located in the North-Eastern lobe where star formation is the most prominent. \section{Summary}\label{sec:summary} We have presented the results of a quantitative spectroscopic analysis of ten O-type stars located in the Local Group dwarf galaxies IC~1613, WLM and NGC~3109. These galaxies have a sub-SMC metallicity based on their oxygen content. \par We derived the wind and atmosphere parameters by adjusting {\sc fastwind} models to the observed line profiles. We derived the fundamental stellar properties (including ages and initial masses) from comparison with evolutionary tracks. \par We used our results to investigate the effective temperature versus spectral type calibration at (sub-)SMC metallicity. We presented both weighted and unweighted fits to the giants and supergiants, and find no significant offset between a calibration based on SMC data and one based on the full sample of stars in IC1613, WLM and NGC~3109 within the limits imposed by our data quality. \par We discussed the location of the sample stars in the Herzsprung-Russell diagram. None of our stars have initial masses higher than $\simeq 50 M_{\odot}$. \par We presented the modified wind momentum versus luminosity diagram. Instead of (sub-)SMC strengths winds, our results indicate stellar winds reminiscent of an LMC metallicity. We discussed the indications that the iron content of the host galaxies may be higher than initially thought, and is possibly SMC-like. While this would lessen the discrepancy with radiation-driven wind theory, the stellar winds of the stars in our sample remain significantly too strong for their metallicity. UV observations of the stars are needed to firmly constrain the wind properties and investigate the effect of wind clumping and the potential presence of an additional wind driving mechanism. \bibliographystyle{aa}
\section{Introduction} Recently, there have been important advances in superconducting large-scale or power applications, partly thanks to the development of ReBCO coated conductors\footnote{ReBCO stands for $Re$Ba$_2$Cu$_3$O$_{7-x}$, where $Re$ is a rare earth, typically Y, Gd or Sm.} and the maturity of MgB$_2$ wires. An important issue in these applications is the electromagnetic design, implying quantities such as the AC loss and, for magnets, the magnetic field quality. Several problems require extensive numerical computations \cite{reviewac}, such as coated conductor magnets containing thousands of turns \cite{neighbour} or 3-dimensional (3D) problems such as wires or cables \cite{nii12SST,amemiya14SST,grilli13Cry,zermeno13SST,stenvall14SST} or motors and generators \cite{masson13IES,zermeno14SST}, among others. Therefore, fast and efficient but still accurate numerical methods for complex situations are required. This work is intended to develop a general variational formalism, including 3-dimensional situations, and apply it to numerically calculate a complex system like magnets with thousands of turns. In the past, there have been significant efforts to compute 3D situations by directly solving a master partial differential equation using finite element methods (FEM) \cite{reviewac}, although for all cases the spatial discretization is relatively coarse, due to the required long computing time. Most of the FEM methods (and the totality of those used in commercial software) need to set the boundary conditions far away from the sample, requiring a high number of elements in the air and increasing the computing time. In order to overcome this problem, the coupled boundary-element/finite-element method (BEM-FEM) was developed \cite{russenschuck99rep,kurz02IES} and showed good results to model generated magnetic fields with superconducting magnets with a ferromagnetic yoke \cite{kurz00IES,hagen12IES}. However, this method has not been applied to electromagnetic time-evolution problems, such as relaxation effects in the superconductor and AC loss. In addition, integral formulations of the ${\bf T}$ current potential have also been shown to reduce the computation region to the superconductor \cite{nii12SST}, although this feature is only applicable for thin films (including surfaces with 3D bending in complicated structures like Roebel cables). Calculations based on variational principles have been shown to be time-efficient, due to both optimized numerical routines and reducing the computation volume to the superconducting region(s) \cite{roebelcomp}. Although it has been shown that for mathematically 2D problems (infinitely long or cylindrical shapes or flat thin films) the region of study can be restricted to the sample volume \cite{prigozhin96JCP,prigozhin97IES,prigozhin98JCP,sanchez01PRB,navau01PRB,HacIacinphase,pancaketheo}, there has not been published any variational principle with this property in 3D. Nevertheless, there have been pioneering contributions to 3D variational principles in the ${\bf H}$ formulation, which requires setting boundary conditions far away from the sample; and thence, requiring elements in the air \cite{bossavit94IEM,badia02PRB,badia12SST}. For all formulations developed up to present, variational principles have limitations to describe ferromagnetic materials interacting with the superconductor. These require to be linear, with either arbitrary \cite{pancakefm} or infinite \cite{sanchez10APLb} permeability. For simplicity, this article regards only the situations with no ferromagnetic materials. An additional feature of the general 3D problem is to determine a realistic ${\bf E}({\bf J})$ relation of the superconductor (where ${\bf E}$ is the electrical field and ${\bf J}$ is the current density) for ${\bf J}$ with a component in the magnetic field direction (flux cutting situation), which causes non-parallel ${\bf E}$ and ${\bf J}$. Although there have been interesting theoretical \cite{clem11PRB,badia09PRB} and experimental \cite{clem11SST} works on this issue, the ${\bf E}({\bf J})$ relation with non-parallel ${\bf E}$ and ${\bf J}$ remains mostly unknown. The present work does not investigate this problem, allowing any ${\bf E}({\bf J})$ relation as input. Independently to the development of numerical models, the computation of the AC loss in coated conductor coils have been an active field of study, by either using variational principles \cite{pancaketheo,pancakefm,pardo12SSTb,neighbour,prigozhin11SST} or solving differential equations by FEM \cite{grilli07SST,ainslie11SST,zermeno11IES,zhangM12APL,zhangM14SST,guC13IES,gomory13IES,vetrella14IES}. However, most of the works only regard single pancake coils (or pancakes) or stacks of few pancakes. In \cite{pardo12SSTb,zhangM14SST}, stacks of many pancakes have been studied but with few turns in each pancake. Nevertheless, \cite{neighbour} calculated the AC loss in a magnet-size coil of 4000 turns (200 turns per pancake), although for the sharp ${\bf E}({\bf J})$ relation of the critical state (figure \ref{f.EJCSM}) with constant $J_c$. However, magnet design (and other power applications) require magnetic field-dependent $J_c$ and smooth ${\bf E}({\bf J})$ relation, the latter being essential to investigate relaxation effects. In this article, we obtain a variational principle for 3D bodies that restrict the calculation volume to the sample (section \ref{s.func}). In that section, we also regard infinitely long or cylindrical symmetries. Afterwards (section \ref{s.nummeth}), we detail the numerical method for circular coils to minimize the functional from the variational principle obtained in the previous section. In section \ref{s.expmeth}, we present the measurement technique used to obtain the input data (the critical current density $J_c$ as a function of the magnetic field and its orientation, and the power-law exponent $n$ of the power-law $E(J)$ relation) and the AC loss, which is compared to the measurements. The next step is to benchmark the numerical method by comparing the calculations to the strip formulas and the experiments on coils made of a few pancakes (section \ref{s.valid}). Once the model is tested, we investigate a magnet-size coil (20 pancakes of 200 turns, totaling 4000 turns) for a smooth $E(J)$ relation and a magnetic field-dependent $J_c$ and discuss the main features of the electromagnetic response, regarding relaxation after the application of a DC input current, cyclic input current, and the effect of magnetization currents on the generated magnetic field (section \ref{s.magnet}). Finally, we present our conclusion (section \ref{s.conc}). \section{Variational principle for 3D bodies} \label{s.func} In the following, we present the 3D variational principle for any material with non-linear ${\bf E}({\bf J})$ relation, such as superconductors. The nathematical method is based on calculating the current density ${\bf J}$ and scalar potential $\phi$ (or magnetic field ${\bf H}$) by minimizing a certain functional. First, we obtain the functional in the ${\bf H}$ formulation and afterwards in the ${\bf J}-\phi$ one. In order to give a certain name, we call the variational principle and the numerical method to solve it presented in this article as Minimum Electro-Magnetic Entropy Production (MEMEP), since the solution minimizes the entropy production due to the electromagnetic fields, as discussed in section \ref{s.thermo}. Actually, there have been several important contributions to the field. First, Bossavit found the 3D functional for the ${\bf H}$ formulation for any ${\bf E}({\bf J})$ relation, including the multi-valued ${\bf E}({\bf J})$ relation of the critical-state model \cite{bossavit94IEM}. However, some key steps in the deduction are omitted in his article. Later on, Badia and Lopez provided a physical insight of the functional and applied the Euler-Lagrange formalism \cite{badia12SST}, although this mathematical framework is strictly only valid for smooth ${\bf E}({\bf J})$ relations. In addition, the ${\bf H}$ formulation has the handicap that the boundary conditions for general sample shapes need to be set far away from the sample, requiring unnecessary elements in the air. Prigozhin introduced the ${\bf J}$ formulation, where the volume of study is reduced to the sample volume, although only for mathematically two-dimensional shapes \cite{prigozhin96JCP,prigozhin97IES,prigozhin98JCP}. This section not only presents a comprehensive deduction of the ${\bf H}$ formulation directly from Maxwell equations with a material ${\bf E}({\bf J})$ relation but also introduces the ${\bf J}-\phi$ formulation for the general 3D case. \subsection{${\bf H}$ formulation} Our goal is to find the functional that by minimizing it, we obtain the solution for ${\bf H}$ (in general, the solution of ${\bf H}$ corresponds to an extreme: minimum, maximum or saddle). Its Euler-Lagrange equations (see \ref{s.EL} on the Euler-Lagrange equations of a functional) should be Faraday's law for a certain ${\bf E}({\bf J})$ relation \begin{equation} \label{faraday} \mu_0{\dot {\bf H}}+\nabla\times{\bf E}({\bf J})=0, \end{equation} where we assume that there are no magnetic materials, ${\bf B}=\mu_0{\bf H}$, and ${\dot {\bf H}}\equiv\partial{\bf H}/\partial t$, with $t$ being the time. Using $\nabla\times{\bf H}={\bf J}$, which corresponds to Ampere's law with negligible displacement current\footnote{The influence of the displacement current, $\partial {\bf D}/\partial t$, on the current distribution in closed current loops (or multi-turn coils) is negligible for conductor lengths, $l$, much shorter than the radiation wavelength $\lambda$ \cite{johnson}. This can be regarded as a rule of thumb for any situation, since in magneto-statics the current always forms a closed loop. Setting a stricter criterion than for antenna design, $l<\lambda/100$ instead of $l<\lambda/10$ (section 5-1 of \cite{johnson}), the displacement current does not influence the current density for frequencies up to around 3 MHz and 3 kHz for conductor lengths of 1 m and 1 km, respectively. Even when the current density in the conductor is not influenced by the displacement current, there will still be a certain small radiation power loss due to the oscillating magnetic dipole moment. This loss is $P_r=\sqrt{\mu_0/\epsilon_0}[4\pi^3N^2A^2I_m^2/(3\lambda^4)]$; where $N$ is the number of turns in the coil, $A$ is the area of one turn, and $I_m$ is the current amplitude (equation (5-5) of \cite{johnson}). However, this contribution is typically negligible compared to the non-linear Joule AC loss in superconductors.}, we obtain the differential equation for ${\bf H}$, \begin{equation} \label{faradayH} \mu_0{\dot{\bf H}}+\nabla\times{\bf E}(\nabla\times{\bf H})=0. \end{equation} We also assume a layered discretization of time. That is, we approximate the time derivatives as ${\dot {\bf H}}\approx\Delta{\bf H}/\Delta t$ with the same time interval for all positions and $\Delta{\bf H}$ is the variation of ${\bf H}$ between two time layers, for instance $t_0$ and $t_0+\Delta t$. Then, Faraday's law becomes \begin{equation} \label{faradayDH} \mu_0\frac{\Delta{\bf H}}{\Delta t}+\nabla\times{\bf E}(\nabla\times{\bf H}_0+\nabla\times\Delta{\bf H})=0, \end{equation} where ${\bf H}_0$ is the magnetic field at the beginning of the interval, $t_0$, while the time at the end is $t_0+\Delta t$. Similarly, the functional and the Euler-Lagrange equations take the form \begin{equation} \label{LHgen} L=\int_V{\rm d}^3{\bf r}f({\bf r},\Delta H_i({\bf r}),\partial_j\Delta H_i({\bf r})) \end{equation} and \begin{equation} \label{ELH} \frac{\partial f}{\partial \Delta H_i}-\sum_{j=1}^3\partial_j\left [ \frac{\partial f}{\partial(\partial_j\Delta H_i)} \right ]=0, \end{equation} respectively, where $V$ is any 3D volume, ${\bf r}=(x_1,x_2,x_3)\equiv(x,y,z)$ and $i\in\{x,y,z\}$. Below, we omit the upper limit of the sums, with the understanding that it is 3. Equations (\ref{faradayH}) and (\ref{ELH}) are conveniently separated into two terms, one depending only on $\Delta H_i$ and another depending on its spacial partial derivatives. Using that $\nabla\times\Delta{\bf H}=\sum_{kji}\epsilon_{kji}\partial_j\Delta H_i{\bf e}_k$, where $\epsilon_{kji}$ is the antisymmetric Levi-Civita symbol and ${\bf e}_k$ is the unit vector in direction $k$, equation (\ref{ELH}) becomes \begin{equation} \label{ELHJ} \frac{\partial f}{\partial H_i}+\sum_{jk}\epsilon_{ijk}\partial_j\left [ \frac{\partial f}{\partial J_k} \right ]=0. \end{equation} Then, if the functional is of the form \begin{eqnarray} L & = & \int_V{\rm d}^3{\bf r}f(\Delta H_i({\bf r}),\partial_j\Delta H_i({\bf r})) \nonumber \\ & = & \int_V{\rm d}^3{\bf r} \left[{ \frac{1}{2}\mu_0\frac{(\Delta {\bf H})^2}{\Delta t}+U({\bf J}_0+\Delta {\bf J}) }\right], \label{fHgen} \end{eqnarray} where ${\bf J}_0=\nabla\times{\bf H}_0$ is the current density at the beginning of the time layer, $\Delta{\bf J}=\nabla\times\Delta {\bf H}$ is the variation of ${\bf J}$ at the end of the time layer, and $U({\bf J})$ is a function such that ${\bf E}=\nabla_{\bf J}U$, being $\nabla_{\bf J}$ the gradient with respect to ${\bf J}$, the Euler-Lagrange equation (\ref{ELHJ}) becomes Faraday's equation, as expressed in (\ref{faradayDH}). Next, we show that for any physical ${\bf E}({\bf J})$ relation there exists a single scalar function defined as \begin{equation} \label{UJ} U({\bf J})=\int_0^{\bf J}{\bf E}({\bf J'})\cdot{\rm d}{\bf J'}. \end{equation} First, $\nabla_{\bf J}\times{\bf E}$ follows \begin{equation} \nabla_{\bf J}\times{\bf E}({\bf J})=\sum_{ijk}\epsilon_{ijk}\frac{\partial E_k({\bf J})}{\partial J_j}{\bf e}_i=\sum_{ijk}\epsilon_{ijk}\rho_{kj}({\bf J}){\bf e}_i. \end{equation} Second, from thermodynamical principles, it can be shown that the differential resistivity matrix $\tilde\rho$, with components $\rho_{kj}$, is symmetric and positive definite \cite{badia12SST}. As a consequence, $\sum_{jk}\epsilon_{ijk}\rho_{kj}=0$, and thence \begin{equation} \nabla_{\bf J}\times{\bf E}=0. \end{equation} Therefore, from Stokes' theorem it follows that $\oint{\bf E}({\bf J}')\cdot{\rm d}{\bf J}'=0$; and hence the scalar function $U({\bf J})$ from equation (\ref{UJ}) is well defined because it does not depend on the integration path. In addition Bossavit found that the $U({\bf J})$ function above can also be applied to the multi-valued ${\bf E}({\bf J})$ relation of the critical state model \cite{bossavit94IEM} (see sketch of the ${\bf E}({\bf J})$ relation for isotropic cases, ${\bf E}\parallel{\bf J}$, in figure \ref{f.EJCSM}). \begin{figure}[tbp] \begin{center} \includegraphics[width=8cm]{fig1.eps}% \caption{\label{f.EJCSM} Qualitative features of the ${\bf E}({\bf J})$ relations for the isotropic power-law ${\bf E}=E_c(|{\bf J|}/J_c)^n({\bf J}/|{\bf J}|)$ and critical-state model (CSM). Bossavit's ${\bf E}({\bf J})$ relation in \cite{bossavit94IEM} actually assumes that the material is linear for $|{\bf E}|$ above a certain threshold.} \end{center} \end{figure} Next, we regard the case that the superconductor is submitted to a certain given applied magnetic field ${\bf H}_a$. In that case there are two contributions to the magnetic field, the applied magnetic field and the magnetic field created by the current density, ${\bf H}_J$; and thence ${\bf H}={\bf H}_J+{\bf H}_a$. The functional (\ref{LHgen}) with the functional density (\ref{fHgen}) becomes \begin{equation} \label{LHa} L=\int_V{\rm d}^3{\bf r}\left [ \frac{1}{2}\mu_0\frac{(\Delta{\bf H}_J)^2}{\Delta t} + \mu_0\Delta{\bf H}_a\frac{\Delta{\bf H}_J}{\Delta t} + U({\bf J}_0+\Delta{\bf J}) \right ], \end{equation} where we dropped the term $(1/2)\mu_0(\Delta{\bf H}_a)^2$ because it does not depend on the minimization variable $\Delta{\bf H}_J$. One possibility to impose a current constraint, that is the total transport current $I$ is given, is to construct an augmented functional with a Lagrange multiplier $\lambda$, \begin{equation} L_a=L+\lambda\left [ \int_{S_1}({\bf J}_0+\Delta{\bf J})\cdot{\rm d}{\bf s}-I \right ]^2=L+\lambda\left [ \oint_{\partial S_1}({\bf H}_0+\Delta{\bf H})\cdot{\rm d}{\bf l}-I \right ]^2, \end{equation} where $L$ is the functional in (\ref{LHa}), $S_1$ is any cross-section where the net current is $I$, $\partial S_1$ is its boundary, ${\rm d}{\bf s}$ is the surface differential and ${\rm d}{\bf l}$ is the path differential. Since we are using that ${\bf J}=\nabla\times{\bf H}$, this implies $\nabla\cdot{\bf J}=0$, and thence there is current continuity in the whole body. For multiply connected superconductors, such as multi-tape cables or coils, one can add as many additional terms in the functional as current constraints. In the minimization of the augmented functional above, the Lagrange multiplier $\lambda$ should be treated as an independent variable. \subsection{${\bf J}-\phi$ formulation} \label{s.Jphi} Next, we deduce an equivalent 3D functional for a formalism depending on the current density ${\bf J}$ and the scalar potential $\phi$. This ${\bf J}-\phi$ formulation allows to greatly reduce the number of variables compared to the $\bf H$ formulation, since the computation volume is restricted to the superconductor. Actually, the ${\bf J}$ formulation also requires computing the scalar potential $\phi$ (or the electric charge density $q$) but the important reduction of variables in the air justifies adding this additional scalar field. In the ${\bf J}$ formulation, we may regard either ${\bf J}$ or the vector potential ${\bf A}$ as state variables. First, we wish to find a fundamental equation for ${\bf A}$ and $\phi$. For this purpose, we take the relation between ${\bf E}$ and the potentials \begin{equation} \label{EAphi} {\bf E}+\dot{\bf A}+\nabla\phi=0. \end{equation} Given a certain ${\bf E}({\bf J})$ relation and using that for Coulomb's gauge, $\nabla\cdot{\bf A}=0$, the Ampere's law $\nabla\times{\bf H}={\bf J}$ becomes\footnote{Again, we neglect the displacement current ($\partial{\bf D}/\partial t\approx 0$) and assume no magnetic materials (${\bf B}=\mu_0{\bf H}$). We also use the definition of vector potential ${\bf B}=\nabla\times{\bf A}$.} $\nabla^2{\bf A}=-\mu_0{\bf J}$. Note that in Coulomb's gauge the scalar potential becomes the electrostatic potential (see Appendix B of \cite{reviewac}). Then, we obtain the system of equations \begin{eqnarray} {\bf E}(-\nabla^2{\bf A}/\mu_0)+\dot{\bf A}+\nabla\phi=0, \label{EAphiCou} \\ \nabla\cdot{\bf A}=0, \label{Cougauge} \end{eqnarray} where the Coulomb's gauge condition is added to the initial relation (\ref{EAphi}). The reason is that (\ref{EAphi}) consists of 3 equations (one for each vector component), while there are 4 fields to be determined (the 3 components of ${\bf A}$ and $\phi$). In addition, the Poisson equation for ${\bf A}$, $\nabla^2{\bf A}=-\mu_0{\bf J}$, does not directly imply the continuity equation for ${\bf J}$, $\nabla\cdot{\bf J}=0$. Explicitly, $\mu_0\nabla\cdot{\bf J}=-\nabla\cdot(\nabla^2{\bf A})=\nabla\cdot[\nabla\times(\nabla\times{\bf A})-\nabla(\nabla\cdot{\bf A})]=-\nabla^2(\nabla\cdot{\bf A})$, and thence $\nabla\cdot{\bf J}=0$ follows for $\nabla\cdot{\bf A}=0$ but not for any gauge. Next, we consider a layered time discretization, so that $\dot{\bf A}({\bf r})\approx\Delta{\bf A}({\bf r})/\Delta t$ with the same $\Delta t$ for any $\bf r$ and where $\Delta {\bf A}({\bf r})$ is the variation of ${\bf A}({\bf r})$ between time $t=t_0$ and $t=t_0+\Delta t$. Then, equations (\ref{EAphiCou}) and (\ref{Cougauge}) become \begin{eqnarray} {\bf E}[-\nabla^2({\bf A}_0+\Delta {\bf A})/\mu_0]+\frac{\Delta{\bf A}}{\Delta t}+\nabla\phi=0, \label{EDAphi} \\ \nabla\cdot({\bf A}_0+\Delta {\bf A})=0, \label{CouDA} \end{eqnarray} where ${\bf A}_0\equiv{\bf A}(t=t_0)$ and we used the fact that equations (\ref{EAphiCou}) and (\ref{Cougauge}) hold for all the previous time layers, and therefore $\nabla\cdot{\bf A}_0=0$. In the following, we present a certain functional and then we proof that its Euler-Lagrange equations are (\ref{EDAphi}) and (\ref{CouDA}). The ansatz of the functional with functional density $f$ is \begin{eqnarray} L & = & \int_V{\rm d}^3{\bf r} f \label{funcAphi} \\ & = & \int_V{\rm d}^3{\bf r} \left[{ \frac{1}{2}\frac{\Delta{\bf A}}{\Delta t}\cdot\Delta{\bf J} + U({\bf J}_0+\Delta{\bf J})+\nabla\phi\cdot({\bf J}_0+\Delta{\bf J}) }\right] \label{fDA} \end{eqnarray} with ${\bf J}_0=-\nabla^2{\bf A}_0/\mu_0$, $\Delta{\bf J}=-\nabla^2\Delta{\bf A}/\mu_0$ and $U({\bf J})$ defined as equation (\ref{UJ}), which follows ${\bf E}=\nabla_{\bf J}U$. Since the functional above depends on the second space derivative of $\Delta{\bf A}$ through $\Delta{\bf J}$, its Euler-Lagrange equations also contain second derivatives (see section \ref{s.EL}) as \begin{eqnarray} \label{ELDA} \frac{\partial f}{\partial \Delta A_i}-\sum_j\partial_j\left [ \frac{\partial f}{\partial(\partial_j\Delta A_i)} \right]+ \sum_{jk} \partial_j\partial_k \left [ \frac{\partial f}{\partial(\partial_j\partial_k\Delta A_i)} \right ]=0 \\ \frac{\partial f}{\partial \phi}-\sum_j\partial_j\left [ \frac{\partial f}{\partial(\partial_j\phi)} \right]=0. \end{eqnarray} The equation for $\phi$ results in \begin{equation} \label{DJ0} \nabla\cdot({\bf J}_0+\Delta{\bf J})=0. \end{equation} Using $\nabla^2{\bf A}=-\mu_0{\bf J}$ and the vector relation $\nabla^2{\bf A}=\nabla\times(\nabla\times{\bf A})-\nabla(\nabla\cdot{\bf A})$, equation (\ref{DJ0}) turns into \begin{equation} \label{D2DA} \nabla^2[\nabla\cdot({\bf A}_0+\Delta{\bf A})]=0. \end{equation} Appliying the Euler-Lagrange equations (\ref{ELDA}) to the functional density we obtain \begin{equation} \label{D2EAphi} \nabla^2\left [{ \frac{\Delta{\bf A}}{\Delta t}+{\bf E}({\bf J}_0+\Delta{\bf J})+\nabla\phi }\right ]=0. \end{equation} In order to deduce the equation above we used that, since $\nabla^2\Delta{\bf A}=-\mu_0\Delta{\bf J}$, then \begin{equation} \frac{\partial f}{\partial(\partial_j\partial_k\Delta A_i)}=-\frac{\partial f}{\partial J_i}\frac{\delta_{jk}}{\mu_0}, \end{equation} where $\delta_{jk}$ is 1 when $j=k$ and 0 otherwise. Therefore, we have obtained that the Euler-Lagrange equations from the functional density (\ref{fDA}), equations (\ref{D2DA}) and (\ref{D2EAphi}), correspond to (\ref{CouDA}) and (\ref{EDAphi}) with a global Laplacian operator. Actually, for a general 3D body the part within the Laplacian of (\ref{D2DA}) and (\ref{D2EAphi}) also vanishes, obtaining equations (\ref{CouDA}) and (\ref{EDAphi}). This is because for any scalar or vector function, for instance $g({\bf r})$ and ${\bf G}({\bf r})$, respectively, the fact that its Laplacian is zero implies $g({\bf r})=0$ and ${\bf G}({\bf r})=0$, as long as those functions are also zero at the boundaries of the volume where their Laplacian vanishes. For finite 3D bodies, the potentials ${\bf A}$ and $\nabla{\phi}$ approach zero at infinity. Since we are neglecting electromagnetic radiation, ${\bf E}$ also vanishes far away from the sample. For the idealization of infinitely long wires or cables transporting a certain net current, the wire or cable actually contains a returning conductor that closes the circuit (see section \ref{s.long} for details). As a consequence, all fields actually vanish at infinity and equations (\ref{D2DA}) and (\ref{D2EAphi}) imply (\ref{CouDA}) and (\ref{EDAphi}), respectively. Then, we have found that the $\Delta{\bf A}$ and $\phi$ that correspond to an extreme of the functional (\ref{funcAphi}) are the solutions of the magnetostatic problem. For the cases that $\nabla \phi$ is given by an external source, such as circular coils or long conductors, it can be proofed that the extreme is a minimum (see \ref{s.convex}). For the general case, it is not clear to the authors whether the extreme is always a minimum. Notice that for a mathematical method that finds the extreme of this functional by changing $\Delta{\bf A}$ and $\phi$, one should take those fields into account for the whole 3D space, or for a volume much larger than the sample volume where it is possible to set known boundary conditions. Although that is feasible by allowing the void to be conductive with a large resistivity, and thence allowing a residual ${\bf J}$ outside the sample; we are departing from the goal of reducing the computational volume. This can be solved as follows. Actually, we can also find the extreme of the functional by using $\Delta{\bf J}$ and $\phi$ instead of $\Delta{\bf A}$ and $\phi$, as long as we keep the condition $\nabla^2\Delta{\bf A}=-\mu_0\Delta{\bf J}$. The solution of $\Delta{\bf A}$ for this Poisson equation is \cite{Jackson} \begin{equation} \label{DAJ} \Delta{\bf A}({\bf r})=\frac{\mu_0}{4\pi}\int_V{\rm d}^3{\bf r} ' \frac{\Delta{\bf J}({\bf r}')}{|{\bf r}-{\bf r}'|}. \end{equation} Notice that from the equation above, $\nabla\cdot\Delta{\bf A}=0$ only when ${\nabla}\cdot\Delta{\bf J}=0$. Then, by using the integral equation above, the functional density of (\ref{fDA}) only depends on $\Delta{\bf J}$. Moreover, if the functional is at an extreme for a certain $\Delta{\bf J}$, it will also be at a extreme with the corresponding $\Delta{\bf A}$ from equation (\ref{DAJ}), and thence the electromagnetic quantities follow equations (\ref{EDAphi}) and (\ref{CouDA}), as well as $\nabla\cdot{\bf J}=0$. Now, the boundary conditions for $\Delta{\bf J}$ can be directly set on the sample surface, and thence the integration volume in the functional ({\ref{funcAphi}) can be restricted to the sample volume. However, the $\Delta{\bf A}$ that $\Delta{\bf J}$ generates extends to the whole 3D space. In case that the sample is submitted to a given applied magnetic field ${\bf H}_a=\nabla\times{\bf A}_a/\mu_0$, where ${\bf A}_a$ is the applied vector potential generated by external currents of given magnitude\footnote{The magnetic field generated by any magnetic material may also be reagarded as that generated by an equivalent magnetizaton current density $\nabla\times{\bf M}$, where ${\bf M}$ is the magnetization.} ${\bf J}_a$, the vector potential may be separated into two contributions, one from the current density in the sample ${\bf J}$ (or $\Delta{\bf J}$) and one from the applied field conribution, and thence ${\bf A}={\bf A}_J+{\bf A}_a$ and $\Delta{\bf A}=\Delta{\bf A}_J+\Delta{\bf A}_a$. For this case, one should use an expression for ${\bf A}_a$ that follows $\nabla\cdot{\bf A}_a=0$. Then, the functional becomes \begin{equation} \label{LDAa} L=\int_V{\rm d}^3{\bf r} \left [ { \frac{1}{2}\frac{\Delta{\bf A}_J}{\Delta t}\cdot\Delta{\bf J}+\frac{\Delta{\bf A}_a}{\Delta t}\cdot\Delta{\bf J}+U({\bf J}_0+\Delta{\bf J})+\nabla\phi\cdot({\bf J}_0+\Delta{\bf J}) } \right ], \end{equation} where $\Delta{\bf A}_J$ is related to $\Delta{\bf J}$ by equation (\ref{DAJ}) and we have used that $\int_V{\rm d}^3{\bf r}\Delta{\bf A}_J\cdot\Delta{\bf J}_a=\int_V{\rm d}^3{\bf r}\Delta{\bf J}\cdot\Delta{\bf A}_a$. In the functional above, we have also dropped the terms with $\Delta{\bf A}_a\cdot\Delta{\bf J}_a$ and $\nabla\phi\cdot{\bf J}_a$, since these quantities are fixed (note that in the term with $\nabla\phi\cdot{\bf J}_a$, $\phi$ refers to the scalar potential in the region where ${\bf J}_a$ is flowing). The boundary conditions for $\Delta{\bf J}$ are the following. For finite 3D samples under an applied magnetic field, one may simply impose that the current does not flow outwards from the sample, and thence $\Delta{\bf J}\cdot{\bf e}_n=0$, where ${\bf e}_n$ is the unit vector perpendicular to the surface. For transposed infinitely long wires and cables, it is necessary to take a periodicity condition into account, in order to reduce the problem to one transposition length or a fraction of it. In the case that there is a certain given transport current $I$, an additonal constraint on $\Delta{\bf J}$ should be imposed, as follows. One option is to set the current constraints is by an augmented functional, such as \begin{equation} L_a=L+\lambda\left [ \int_{S_1}({\bf J}_0+\Delta{\bf J})\cdot{\rm d}{\bf s}-I \right ]^2, \end{equation} where $L$ is the functional of (\ref{LDAa}), $S_1$ is any cross-section that transports the current $I$, ${\rm d}{\bf s}$ is the surface differential, and $\lambda$ is a Lagrange multiplier that has to be treated as an independent variable. However, our numerical method for minimization for coils (section \ref{s.genmeth}) takes implicitly the current constraint, and thence an augmented functional is not necessary. In principle, for general 3D problems one may have difficulties setting the boundary conditions for $\phi$ (or $\nabla\phi$) on the sample surface. If this problem arises, it may be solved by using that, since we are using Coulomb's gauge, $\phi$ is the electrostatic potential, and thence it is related to the surface and volume charge densities, $\sigma$ and $q$ respectively, as \begin{equation} \phi({\bf r})=\frac{1}{4\pi\epsilon_0} \left [ { \int_V{\rm d}^3{\bf r} ' \frac{q({\bf r}')}{|{\bf r}-{\bf r}'|} + \oint_{\partial V} {\rm d}{s}\frac{\sigma({\bf r}')}{|{\bf r}-{\bf r}'|} } \right ], \end{equation} where $\partial V$ is the volume surface and ${\rm d}{s}$ is its differential. With this approach, the variables are $\Delta{\bf J}$, $q$ and $\sigma$, which are all constricted within the sample volume. \subsection{Thermodynamical interpretation} \label{s.thermo} Previously, Badia and Lopez provided a thermodynamical interpretation for situations close to the critical-state model using the ${\bf H}$ formulation \cite{badia02PRB,badia12SST}. In the following, the extend and detail the analysis for any ${\bf E}({\bf J})$ relation, also for the ${\bf J}-\phi$ formulation. First, we outline a description on the energetic meaning of the several terms of the functionals (\ref{fHgen}) and (\ref{fDA}) in the ${\bf H}$ and ${\bf J}-\phi$ formulations, respectively. For this purpose, we take into account that for the limit of $\Delta t\to 0$, we obtain $U({\bf J}_0+\Delta{\bf J})-U({\bf J}_0)\approx{\bf E}_0\cdot\Delta{\bf J}$, where ${\bf E}={\bf E}(t_0)$. Then, the functionals become, save a constant term with $U({\bf J}_0)$, \begin{equation} \label{LDHDt} L\approx\frac{1}{\Delta t}\int_V{\rm d}^3{\bf r} \left[{ \frac{1}{2}\mu_0(\Delta{\bf H})^2+\Delta t\ {\bf E}_0\cdot\Delta{\bf J} }\right] \end{equation} for the ${\bf H}$ formalism and \begin{equation} L\approx\frac{1}{\Delta t}\int_V{\rm d}^3{\bf r} \left[{ \frac{1}{2}\Delta{\bf A}\cdot\Delta{\bf J}+\Delta t\ {\bf E}_0\cdot\Delta{\bf J}+\Delta t\ \nabla\phi\cdot({\bf J}_0+\Delta{\bf J}) }\right] \label{LDJDt} \end{equation} for the ${\bf J}-\phi$ one. The first term in (\ref{LDHDt}) is the magnetic energy of the magnetic field variation $\Delta{\bf H}$ ignoring the interaction with the pre-existing magnetic field ${\bf H}_0$, while the second term is twice the heat generated during the time interval $\Delta t$ due to the onset of $\Delta{\bf J}=\nabla\times\Delta{\bf H}$ (here we use that for the first Taylor approximation $\Delta{\bf J}$ increases linearly with time). Regarding the functional in the ${\bf J}-\phi$ formulation, the second term is identical to the ${\bf H}$ formulation and the first term is, similarly, the magnetic energy of $\Delta J$ ignoring the presence of the pre-existing current density ${\bf J}_0$. The third term is twice the energy transferred to the electrostatic system, as long as $\bf A$ is in Coulomb's gauge; and thence $\phi$ is the electrostatic potential. Although there are strong similarities with the Lagrangian formalism of classical mechanics, as Badia and Lopez showed in \cite{badia12SST}, this analogy is incomplete because the first term is not the total energy of the system. Therefore, it is not clear that the functionals in our system can be interpreted as Lagrangians in the classical sense. In the following, we investigate the resulting functional from minimizing the magnetic variable, $\Delta{\bf H}$ or $\Delta{\bf J}$, and its interpretation. Regarding the ${\bf J}-\phi$ formulation, the system follows the Euler-Lagrange equations of the functional, which correspond to (\ref{EAphi}) and (\ref{DJ0}). By using these equations and the relation $\nabla\phi\cdot{\bf J}=\nabla\cdot(\phi{\bf J})-\phi\nabla\cdot{\bf J}$, the minimized functional becomes \begin{eqnarray} L_{\rm min} & = & \int_V {\rm d}^3{\bf r} \left[{ \frac{1}{2} {\bf E}_0\cdot\Delta{\bf J} + \frac{1}{2} \nabla\cdot(\phi{\bf J}) }\right] \nonumber \\ & = & \frac{1}{2} \int_V {\rm d}^3{\bf r} {\bf E}_0\cdot\Delta{\bf J} + \frac{1}{2} \oint_{\partial V} {\rm d}{\bf s}\cdot{\bf J}\phi, \label{Lminfull} \end{eqnarray} where in the last step we used Stokes Theorem. In equation (\ref{Lminfull}) above, the first term is the average heat rate generation due to the onset of $\Delta{\bf J}$ during the time interval $\Delta t$ and the second one is the energy rate flowing outwards the sample. If we include all sources of current in the system, as we have done in the previous sections, the second term drops. Then, \begin{equation} \label{Lmin} L_{\rm min} \approx \int_V {\rm d}^3{\bf r} \frac{1}{2} {\bf E}_0\cdot\Delta{\bf J}. \end{equation} Following a similar argument, the functional for the ${\bf H}$ formalism in (\ref{LDHDt}) results also in (\ref{Lmin}). Therefore, the electromagnetic solution obtained from minimizing both functionals is identical. Notice that $L_{\rm min}$ from the equation above is the average rate of heat generation between time $t_0$ and $t_0+\Delta t$ due to $\Delta{\bf J}$. This is because for a superconductor ${\bf E}\cdot{\bf J}$ is the density of local heat rate generation, as justified in \cite{reviewac}. The total rate of heat generation is \begin{eqnarray} {\dot Q} & = & \int_V {\rm d}^3{\bf r} \left[{ \frac{1}{2} {\bf E}_0\cdot\Delta{\bf J} + {\bf E}_0\cdot{\bf J}_0 }\right] \nonumber \\ & = & L_{\rm min}+\int_V {\rm d}^3{\bf r} \ {\bf E}_0\cdot{\bf J}_0, \label{Qprod} \end{eqnarray} resulting in a heat variation \begin{equation} \delta Q=L_{\rm min}\delta t+\delta t\int_V {\rm d}^3{\bf r} \ {\bf E}_0\cdot{\bf J}_0. \end{equation} Since from the second law of thermodynamics $\delta Q=T\delta S$, where $T$ is the temperature and $\delta S$ is the entropy, then the rate of entropy production is \begin{equation} {\dot S}=\frac{1}{T}\left[{ L_{\rm min}+\int_V {\rm d}^3{\bf r} \ {\bf E}_0\cdot{\bf J}_0 }\right]. \end{equation} Since the second term does not depend on $\Delta{\bf J}$, we have found that for isothermal conditions (as it is usually assumed for purely magnetic modelling) the $\Delta{\bf J}$ that minimizes the functional $L$ also minimizes the rate of entropy production. In addition, the rate of heat production, equation (\ref{Qprod}), is also minimum. This suggests that the functionals from (\ref{LDHDt}) and (\ref{LDJDt}) may correspond to the entropy production, save constant terms. \subsection{Long straight wires and cables} \label{s.long} In this section, we present the modifications to the functional in the ${\bf J}-\phi$ formulation for infinitely long straight wires or cables (referred below as ``conductors") transporting a certain current $I$. Let us take $z$ as the direction that the conductor extends infinitely. Then, the current density and vector potential follow the $z$ direction and do not depend on $z$; and thence $\Delta{\bf J}({\bf r})=\Delta J(x,y){\bf e}_z$ and $\Delta{\bf J}({\bf r})=\Delta J(x,y){\bf e}_z$, where $x$ and $y$ are the other Cartesian components and ${\bf e}_z$ is the unit vector in the $z$ direction. For this case, the magnetic induction ${\bf B}$ is perpendicular to ${\bf J}$, and thence there is no flux cutting. As a result, ${\bf E}$ is parallel to ${\bf J}$ and ${\bf E}({\bf J})=E(J){\bf e}_z$, where $E$ has the same sign as $J$. Then from ${\bf E}({\bf r})=E(x,y){\bf e}_z$ and equation (\ref{EAphi}) follows that $\nabla\phi({\bf r})=\partial_z\phi{\bf e}_z$, which is constant within the conductor (or each tape or filament in multi-tape or multi-filament conductors). In addition, the function for $U$ in (\ref{UJ}) can be simplified as $U(J)=\int_0^J E(J'){\rm d} J'$. Then, the functional in (\ref{LDAa}) becomes \begin{eqnarray} L & = & l\int_S{\rm d}^2{\bf r} f \nonumber \\ & = & l\int_S{\rm d}^2{\bf r} \left[{ \frac{1}{2}\frac{\Delta A_J}{\Delta t}\Delta J+\frac{\Delta A_a}{\Delta t}\Delta J+U(J_0+\Delta J)+\partial_z\phi({\bf J}_0+\Delta{\bf J}) }\right], \label{Linf} \end{eqnarray} where $l$ is the conductor length, $S$ is the superconductor cross-section in the $xy$ plane, and ${\rm d}^2{\bf r}$ is ${\rm d} x{\rm d} y$. Since any conductor transporting a certain current should form a closed circuit, we may consider a returning conductor separated by a certain distance $D$ much larger than both the conductor width and thickness but still much shorter than the conductor length $l$. The functional of this system is \begin{equation} \label{Linfret} L=l\int_{S_+}{\rm d}^2{\bf r} f+l\int_{S_-}{\rm d}^2{\bf r} f, \end{equation} where $f$ is the functional density of equation (\ref{Linf}), $S_+$ is the section of the conductor transporting current $I$ (or ``main" conductor) and $S_-$ is the returning conductor, with transport current $-I$. Next, we pay attention to the ``main" conductor only. The variation of vector potential has two components, $\Delta A=\Delta A_J+\Delta A_{\rm int}$, regarding the contribution from $\Delta J$ in the ``main" conductor and the interaction with the returning one. By direct integration of (\ref{DAJ}) for a wire of length $l$ without adding any additional constant\footnote{Far away from a wire of arbitrary cross-section and arbitrary internal distribution of current, the vector potential is the same as an infinitesimally thin wire.}, $A_{\rm int}\approx-\mu_0I\ln(l/D)/(2\pi)$. It is important to notice that the interaction term $A_{\rm int}$ is constant. Therefore, $\Delta J$ in the ``main" conductor is independent on $\Delta J$ in the returning one, as long as the total current is fixed. As a consequence, the two terms of the functional (\ref{Linfret}) can be minimized independently. For the ``main" wire, the functional turns into \begin{eqnarray} L & = & l\int_{S_+}{\rm d}^2{\bf r} \left[{ \frac{1}{2}\Delta A_J\Delta J+\Delta A_a\Delta J+U(J_0+\Delta J)+\partial_z\phi({\bf J}_0+\Delta{\bf J}) }\right] \nonumber \\ & & -l\frac{\mu_0}{4\pi}(\Delta I)^2\ln\left( \frac{l}{D} \right). \end{eqnarray} For minimization purposes, the last term is constant and could be dropped, as well as the general $l$ factor. \subsection{Axi-symmetric systems and coils} \label{s.cylsym} This section obtains a simplified functional for the ${\bf J}-\phi$ formalism valid for axi-symmetric systems, also regarding multiple connected bodies made of concentric rings. For bodies with axial symmetry, ${\bf J}$ (and $\Delta{\bf J}$) follow the angular direction and do not depend on the angular coordinate $\varphi$. Therefore, $\Delta{\bf J}({\bf r})=\Delta J(r,z){\bf e}_\varphi$ and $\Delta{\bf A}({\bf r})=\Delta A(r,z){\bf e}_\varphi$, where $r$ and $z$ are the radial and axial components and ${\bf e}_\varphi$ is the unit vector in the angular direction. As for infinitely long conductors, axi-symmetric superconductors do not present flux cutting and ${\bf E}({\bf J})=E(J){\bf e}_\varphi$, where $E$ has the same sign as $J$. Then, the function for $U$ in (\ref{UJ}) becomes simply $U(J)=\int_0^J E(J'){\rm d} J'$. Finally, as a consequence of (\ref{EAphi}) and the axial symmetry, $\nabla\phi(r,z)=(1/r)\partial_\varphi\phi{\bf e}_\varphi$ and $\partial_\varphi\phi$ is constant within each isolated ring (or each turn separately for a coil with axial symmetry). Circular coils may be approximated as a set of concentric rings with given current $I$. The drop of electrostatic potential that drives the current in each turn may be regarded either as a separate voltage source in each turn or, more realistically, as a global source but with turns that break the circular symmetry in only one point in order to connect with its neighbour turn (figure \ref{f.sketchcoil}). \begin{figure}[tbp] \begin{center} \includegraphics[width=6cm]{fig2.eps}% \caption{\label{f.sketchcoil} Closely packed pancake coils can be well approximated as an axi-symmetric problem. The turns are assumed circular, except at a small section where it connects with the following turn. At that point the voltage drop at a certain turn $i$ is defined $\Delta \phi_i$.} \end{center} \end{figure} Therefore, the functional of (\ref{LDAa}) becomes \begin{eqnarray} \label{Lcyl} L=2\pi\sum_{i=1}^{n_t} \bigg( & \int_{S_i} & {\rm d}{s} \ r\left [{ \frac{1}{2}\frac{\Delta A_J}{\Delta t}\Delta J+\frac{\Delta A_a}{\Delta t}\Delta J+U(J_0+\Delta J) }\right ] \nonumber \\ & + & \partial_\varphi\phi_i\int_{S_i}{\rm d}s({\bf J}_0+\Delta{\bf J}) \bigg), \end{eqnarray} where $n_t$ is the number of simply connected regions (or number of turns in coils), ${\rm d}s$ is ${\rm d}r{\rm d}z$, $S_i$ is the surface cross-section of region $i$, and $\partial_\varphi\phi_i$ is $\partial_\varphi\phi$ at the same region $i$. In principle, the current constraint may be set by Lagrange multipliers, as described in section (\ref{s.Jphi}). However, our minimization process maintains a constant net current, only considering variations of $\Delta J$ that do not modify the net current. Therefore, the last term in (\ref{Lcyl}) becomes $\partial_\varphi\phi_i I_i$, where $I_i$ is the net current. This term does not depend on the particular distribution of $\Delta J$, and thence it may be dropped from the functional, resulting in \begin{equation} \label{Lcyl} L=2\pi \int_{S} {\rm d}{s} \ r\left [{ \frac{1}{2}\frac{\Delta A_J}{\Delta t}\Delta J+\frac{\Delta A_a}{\Delta t}\Delta J+U(J_0+\Delta J) }\right ], \end{equation} where all surface integrals are merged into one in order to simplify the notation. \subsection{Magnetic field-dependent or position-dependent ${\bf E}({\bf J})$ relation} The electromagnetic problem can be solved by minimizing the functional in (\ref{LDAa}) also for magnetic field-dependent and position-dependent ${\bf E}({\bf J})$ relations. In practice, the ${\bf E}({\bf J})$ relation depends on the magnetic field\footnote{Actually, ${\bf B}$ is the magnetic flux density. However, since we assume no magnetic materials, ${\bf B}=\mu_0{\bf H}$, and thence $\bf B$ and $\bf H$ play the same physical role, except of a constant. Therefore, in this article we name both as ``magnetic field" for simplicity.} $\bf B$; such as an isotropic power-law relation ${\bf E}=E_c(|{\bf J}|/J_c)^n({\bf J}/|{\bf J}|)$ with magnetic field-dependent parameters $J_c({\bf B})$ and $n({\bf B})$, and constant $E_c$. This case can be solved iteratively as follows. First, ${\bf B}$ is taken as that at $t=t_0$. Then, $\Delta{\bf J}$ and $\phi$ are solved by minimizing (\ref{LDAa}). Afterwards, ${\bf B}$ is calculated again and the process is repeated until the difference in $\Delta {\bf J}$ and $\phi$ between two iterations is below a certain tolerance (more details in section \ref{s.EJBmeth}). For a position-dependent ${\bf E}({\bf J})$ relation, ${\bf E}({\bf J},{\bf r})$, one simply has to take the corresponding position-dependent $U({\bf J},{\bf r})=\int_0^{\bf J}{\rm d}{\bf J}'\cdot{\bf E}({\bf J}',{\bf r})$. An example of these ${\bf E}({\bf J},{\bf r})$ relations are isotropic power laws with position-dependent $J_c$ or $n$, caused by are either non-uniform material properties \cite{jiangZ07PhC,solovyov13SST} or thickness variations in thin films \cite{tsukamoto05SST}. \section{Numerical method for coils} \label{s.nummeth} In this section, we present the details of the numerical method to obtain the time evolution of $\bf J$ in a superconducting coil, and from this result calculate the rest of the electromagnetic parameters, such as the generated magnetic field, the critical current, the AC loss, and the coil voltage. In short, given the current density for a certain time $t_0$, ${\bf J}_0$, the method finds the current density ${\bf J}={\bf J}_0+\Delta{\bf J}$ for a time $t=t_0+\Delta t$. This is done by minimizing the functional (\ref{LDAa}) by keeping the total transport current in the coil constant. We also assume an axi-symmetric symmetry of the coil, as outlined in section \ref{s.cylsym}. However, the method presented here can also be applied to long conductors by taking a closed ring loop of the conductor and setting an average radius much larger than the conductor width and thickness. \subsection{$E(J)$ relation} Although the numerical method presented here is valid for any $E(J)$ relation, the results in this work are for a power-law expression as \cite{brandt95RPP,plummer87IEM,rhyner93PhC} \begin{equation} \label{EJ} E(J)=E_c\left( \frac{|J|}{J_c} \right)^n \frac{J}{|J|}, \end{equation} where the parameters are the critical current density $J_c$, the power-law exponent $n$ and the voltage criterion for the critical current density $E_c$. The relation above is generally valid for superconductors with $J$ close to the critical current density \cite{brandt95RPP}. With this ${\bf E}({\bf J})$ relation, the function $U({\bf J})$ in equation (\ref{UJ}) becomes \begin{equation} U(J)= \frac{1}{n+1} E_cJ_c\left( \frac{|J|}{J_c} \right)^{n+1}, \end{equation} where $J$ is the (only) axial component of ${\bf J}$. In general, the parameters $J_c$ and $n$ of the power law depend of the magnetic field $\bf B$. In this article, we use $J_c({\bf B})$ and $n({\bf B})$ dependencies extracted from measurements (see sections \ref{s.Jcn} and \ref{s.JcBmagnet}). In a similar way, the numerical method is prepared to take position-dependent parameters, $J_c(r,z)$ and $n(r,z)$, into account; although we do not present results in this work. \subsection{Discretization} In order minimize the functional and find the solution of the current density, we divide the entire cross-section into $N$ elements where we assume uniform current density. The computations in this article are done with a uniform mesh discretization with elements of rectangular cross-section. However, the method is also valid for elements of any cross-section shape, such as triangular ones, at least when the cross-sectional area of all elements is the same \cite{ruuskanen14prp}. The formalism below is written taking this into account. The reason to use a uniform mesh is to minimze the RAM memory (see the last paragraph of this section for details). Given a certain element $i$, we define $I_i(t)$ as the current in that element at a time $t$. Then, the current density at that element is $J_i=I_i/S_i$, where $S_i$ is the cross-section of element $i$. In consistency with the notation in section (\ref{s.Jphi}), we denote $I_{0,i}=J_{0,i}/S_i$ and $\Delta I_i=\Delta J_i/S_i$, where $I_{0,i}$ and $\Delta I_i$ are the current in element $i$ at the previous time, $t_0$, where $J$ is solved and $\Delta I_i$ is the change in current in element $i$ when increasing the time by $\Delta t$. We also define the average magnetic flux in element $i$ cross-section as \begin{equation} F_i=\frac{2\pi}{S_i}\int_{S_i}{\rm d}s\ rA(r,z). \end{equation} Similarly, quantities $\Delta F_{J,i}$ and $\Delta F_{a,i}$ are defined as $F_i$ but replacing $A$ by $\Delta A_J$ and $\Delta A_a$, respectively. With our discretization, $\Delta F_i$ from the definition above and equation (\ref{DAJ}) becomes \begin{equation} \Delta F_{J,i}=\sum_iC_{ij}\Delta I_i, \end{equation} where the sum is done for all elements, $1 \le i \le N$, and the constant terms $C_{ij}$ are \begin{equation} \label{cij} C_{ij}=\frac{2\pi}{S_iS_j}\int_{S_i}{\rm d} s \int_{S_j}{\rm d} s' ra_{\rm loop}(r,r',z-z'). \end{equation} In the equation above, ${\bf r}$ and ${\bf r}'$ are 3D vector positions, while $r$ and $r'$ are the radial components only, and $a_{\rm loop}$ is the vector potential generated by a circular loop per unit current in the loop with radius $r'$ located at height $z'$. The expression of this function is given by equations (\ref{aloop}) and (\ref{k}) in \ref{s.intmat}. The matrix elements $C_{ij}$ are numerically evaluated, as detailed in \cite{pancaketheo}. Using the $C_{ij}$ matrix, the functional in (\ref{Lcyl}) for our discretization becomes \begin{equation} L=\frac{1}{2\Delta t}\sum_{ij}C_{ij}\Delta I_i\Delta I_j+\frac{1}{\Delta t}\sum_i \Delta F_{a,i}\Delta I_i + 2\pi\sum_ir_iS_iU_i, \end{equation} where $U_i\equiv U[(I_{0,i}+\Delta I_i)/S_i]$. Next, for minimization purposes, we regard a change in $L$ due to a change $\delta I$ of $\Delta I_i$. The resulting change in $L$ is \begin{eqnarray} \delta L_i & = & \frac{1}{\Delta t}(\Delta F_{J,i}+\Delta F_{a,i})\delta I+\frac{1}{2\Delta t}C_{ii}(\delta I)^2 \nonumber \\ & & +2\pi r_iS_i \left [{ U\left( \frac{I_{0,i}+\Delta I_i+\delta I}{S_i} \right) - U_i }\right ]. \end{eqnarray} For the rest of this article, the quantities between braces refer to the vector composed by the value of that quantity among all elements, such as $\{I_i\}$, $\{\Delta F_i\}$ or $\{U_i\}$. Since we take a magnetic field-dependent $J_c$ and $n$ into account, we also need to compute the magnetic field in the superconductor. This is done numerically, as follows. For our discretization, we take the average magnetic field in a given element $i$ cross-section as the relevant quantity to calculate $J_c({\bf B}$) and $n({\bf B})$ in that element, that is \begin{equation} {\bf B}_i\equiv\frac{1}{S_i}\int_{S_i}{\rm d} s {\bf B}({\bf r}). \end{equation} Accordingly, we define ${\bf B}_{J,i}$ and ${\bf B}_{a,i}$ by substituting ${\bf B}$ in the equation above by the magnetic field created by the currents and the applied field, ${\bf B}_J$ and ${\bf B}_a$, respectively. Since ${\bf B}_a$ is usually analytical, it can be straightforwardly calculated. By means of the Biot-Savart law, ${\bf B}_{J,i}$ can be calculated as the sum of the contributions from all elements as \begin{equation} {\bf B}_{J,i}=\sum_i{\bf b}_{ij}I_j \end{equation} with \begin{equation} \label{bij} {\bf b}_{ij}=\frac{1}{S_iS_j}\int_{S_i} {\rm d} s \int_{S_j} {\rm d} s' {\bf b}_{\rm loop}(r,r',z-z'), \end{equation} where ${\bf b}_{\rm loop}$ is the magnetic field created by an infinitely thin circular loop per unit current in the loop, given by equations (\ref{brloop})-(\ref{k}) in \ref{s.intmat}. That appendix also presents the numerical method to evaluate ${\bf b}_{ij}$. For meshes with rectangular elements of identical cross-section, the independent entries of the interaction matrices $C_{ij}$, $b_{r,ij}$ and $b_{z,ij}$ can be greatly reduced comparing to arbitrary non-uniform meshes. For the latter, the each interaction matrix contains $N^2$ independent entries, where $N$ is the total number of elements. For the former, the number of independent entries in the matrices can be reduced as follows. From equations (\ref{cij}) and (\ref{bij}) we can see that, $C_{ij}$, $b_{r,ij}$ and $b_{z,ij}$ only depend on $r_i$, $r_j$ and $z_i-z_j$, where $(r_i,z_i)$ and $(r_j,z_j)$ are the coordinates of the cross-sectional center of elements $i$ and $j$, respectively. For given a $r_i$ and $r_j$, the number of different $z_i-z_j$ is only $n_z(2n_p-1)$, where $n_z$ and $n_p$ are the number of elements in the $z$ direction in each single tape and are the number of pancakes, respectively. Then, the number of independent entries in each interaction matrix is $n_r^2n_z(2n_p-1)$, which is much smaller than the one for the general case $N^2=(n_rn_zn_p)^2$, where $n_r$ is the number of elements in the $r$ direction. As a result, uniform mesh allows a reduction in RAM memory storage or the interaction matrices, which occupies most of the memory storage of the program, by a factor $n_zn_p/(2-1/n_p)$. For $n_z=100$, this reduction is by a factor around 100 and 1000 for 1 and 20 pancakes, respectively. \subsection{General minimization method} \label{s.genmeth} As mentioned above, our minimization method contains the current constraints as a built-in feature, and therefore minimization with Lagrange multipliers is not necessary. The main steps of the minimization process are the following (see algorithm \ref{a.genmeth}). We start with a physical current distribution, $\{I_i\}$, corresponding to a certain time $t_0$, and assign 0 to $\{\Delta I_i\}$. Then, we set the net transport current in each turn by distributing the change in transport current, $\Delta I_{\rm tran}$, uniformly over the cross-section of each turn. Next, we find the induced magnetization currents, as follows. For each turn, we find the element $i_+$ where adding a certain positive value $\delta I=h$ to $\Delta I_{i_+}$ decreases $L$ the most (or increases it the least). In this routine, the value of $h$ sets the tolerance. We continue by finding the element $i_-$ within the same turn where substracting $h$ (or adding $\delta I=-h$) to $\Delta I_{i_-}$ minimizes the most $L$. We do the same operation for all turns, so that we find the pair $i_+$ and $i_-$ that minimizes the most $L$ for the whole coil, taking elements $i_+$ and $i_-$ that belong to the same turn. Thus, we reduce $L$ while keeping the current constraint. Once we have found the optimum pair, we set the new values to the pair, $\Delta I_{i_+}\gets \Delta {I_{i_+}}+h$ and $\Delta I_{i_-}\gets \Delta {I_{i_-}}-h$. Afterwards, we update $\{\Delta F_i\}$ and $\{U_i\}$ accordingly. In this way, we do not need to evaluate these quantities at each evaluation of $\delta L$, only when the change in current is finally set. We continue this process until any pair of elements with $\delta I=+h$ and $\delta I=-h$ will increase $L$ instead of decreasing it. This minimization routine will always obtain the current distribution corresponding to the global minimum of $L$ within the tolerance $h$, as detailed in \cite{HacIacinphase}. \begin{algorithm} \caption{The minimization method in pseudo-code below rapidly calculates the current distribution in the coil by minimizing the functional in (\ref{Lcyl}). More details are provided in the text.}\label{a.genmeth} \begin{algorithmic}[l] \State Set $\Delta I_i \gets 0$ for all $i$; \State Add $\Delta I_{\rm tran}$ distributed uniformly among all elements; \Repeat \State $\delta L\gets 1$; \For{${\rm turn}=1$ to $n_t$} \State Find the element $i_+$ within the present turn \State where {\em adding} $h$ to $\Delta I_{i_+}$ produces the smallest $\delta L_{i_+}$; \State Find the element $i_-$ within the present turn \State where {\em substracting} $h$ from $\Delta I_{i_-}$ produces the smallest $\delta L_{i_-}$; \If{ $\delta L>\delta L_{i_+}+\delta L_{i_-}$ } \State $\delta L\gets \delta L_{i_+}+\delta L_{i_-}$; \State ${i_{+,{\rm min}}}\gets i_+$; \State ${i_{-,{\rm min}}}\gets i_-$; \EndIf \EndFor \If{ $\delta L < 0$ } \State $\Delta I_{i_{+,{\rm min}}}\gets \Delta I_{i_{+,{\rm min}}}+h$; \State $\Delta I_{i_{-,{\rm min}}}\gets \Delta I_{i_{-,{\rm min}}}-h$; \State Update $F_i$ and $U_i$ for all $i$; \EndIf \Until{$\delta L\ge 0$;} \end{algorithmic} \end{algorithm} \subsection{Method for magnetic field-dependent $E(J)$ relation} \label{s.EJBmeth} For a magnetic field-dependent $E(J)$, such as a power-law with $J_c({\bf B})$ and $n({\bf B})$, we use an iterative method as follows (see algorithm \ref{a.JcBmeth}). First, we add the change in transport current $\Delta I_{\rm tran}$ uniformly among each turn. Afterwards, we find the current distribution $\{\Delta I_i\}$ that minimizes the functional while maintaining the value of transport current, as detailed in section \ref{s.genmeth}. In order to avoid oscillations between iterations, we apply a damping factor to $\{\Delta I_i\}$, such as \begin{equation} \Delta I_i\gets \Delta I_{p,i}+(\Delta I_{i}-\Delta I_{p,i})K_d, \end{equation} where $\Delta I_{p,i}$ is the change in current of element $i$ at the previous iteration and $K_d$ is the damping factor. We found an optimum value of $K_d=0.9$ regarding computing time. Afterwards, we calculate both components of the magnetic field in all the elements, $\{B_{r,i}\}$ and $\{B_{z,i}\}$, and update the vectors containing $J_c$ and $n$, $\{J_{c,i}\}$ and $\{n_i\}$. Finally, we need to update $U$ for each element, $\{U_i\}$, as a consequence of the local change of $J_c$ and $n$. We repeat the iterations until the change of $\Delta I_i$ is below the tolerance $h$ for any element $i$. The speed of the algorithm have been increased as follows. We start with an initial $h$ much larger than our final tolerance goal as $h=2^kh^*$, where $h^*$ is our tolerance goal and $k$ is an integer larger than one. Afterwards, we repeat the minimization process but with half the previous $h$, until $h=h^*$, our desired value. We have found an optimum exponent of $k=5$. We choose $h^*=J_{c0}S_{\rm av}/m$, where $J_{c0}=J_c(B=0)$, $S_{\rm av}$ is the average cross-section of the elements, and $m$ is an integer number ranging from 500 to 45000, being the largest values for the lowest current amplitudes. Computations for lower current amplitudes require more strict tolerance $h^*$ because, for the same time step, the average change in the current density decreases with the current amplitude, and thence the same absolute error $h^*$ corresponds to a higher relative error. \begin{algorithm} \caption{The interative method below in pseudo-code obtains the current distribution for an $E(J)$ relation with magnetic field-dependent parameters, such as the critical current density $J_c$ and the power-law exponent $n$.}\label{a.JcBmeth} \begin{algorithmic}[l] \State Set $\Delta I_i \gets 0$ for all $i$; \State Add $\Delta I_{\rm tran}$ distributed uniformly among all elements; \Repeat \State Set $\Delta I_{p,i}\gets \Delta I_{i}$ for all $i$; \State Find $\{\Delta I_i\}$ that minimizes functional \State while keeping the value of transport current; \State Set $\Delta I_i\gets \Delta I_{i,p}+(\Delta I_{i}-\Delta I_{p,i})K$ for all $i$; \State Calculate $\{B_{r,i}\}$ and $\{B_{z,i}\}$; \State Update $\{J_{c,i}\}$ and $\{n_i\}$; \State Update $\{U_{i}\}$; \Until{change in $\Delta I_i$ below tolerance for any $i$;} \State Set $I_i\gets I_i+\Delta I_i$ for all $i$; \end{algorithmic} \end{algorithm} \subsection{AC loss calculation} Once $J$ is known, the instantaneous power loss can be simply evaluated as \begin{equation} P=\int_V{\rm d}^3{\bf r}\ {\bf E}({\bf J})\cdot{\bf J}=2\pi\int_S{\rm d} s\ rE(J)J, \end{equation} since ${\bf E}\cdot{\bf J}$ describes the local heat generation \cite{reviewac}. Thus, the loss per cycle (or heat generated per cycle) is simply \begin{equation} Q=\oint_T{\rm d} t\int_V{\rm d}^3{\bf r}\ {\bf E}\cdot{\bf J}=2\pi\oint_T{\rm d} t\int_S{\rm d} s\ rE(J)J, \end{equation} where $T$ is the period of the external excitation. For any transport current or applied magnetic field with a symmetrical waveform, such as triangular or sinusoidal, the integral can be reduced to half a period. In this article, we assume sinusoidal transport currents, $I(t)=I_m\sin(2\pi\nu t)$, of a given frequency $\nu$. For this case, we calculate the ac loss over the half cycle after the first instant that $I=-I_m$. This is because for conductors submitted to simultaneous alternating current and magnetic field, such as in coils, the loss signal becomes periodic after that instant \cite{HacIacinphase}. Alternatively, the AC loss per cycle could be calculated from the power delivered from the power source \cite{reviewac} \begin{equation} Q=\oint_T{\rm d} t\ v\cdot I, \end{equation} where $v$ is the voltage in the coil, calculated as detailed in section $\ref{s.vphi}$. Notice that $v\cdot I$ is not necessarily the instantaneous power dissipation in the coil, due to inductive effects. In this article, we typically divide the AC cycle into 80 equal time steps for AC loss calculations. \subsection{Implementation and computing times} The numerical implementation is programmed in Fortran 95; although C++, MATLAB or other general-purpose programming languages may be used. The computations in this article have been done with either a table computer with Intel(R) Core(TM) i7-3770K CPU processor and 8 GB RAM or a server with two Intel(R) Xeon(TM) E5645 processors and 48 GB RAM, both presenting similar computing time. The program also executes computations corresponding to different current amplitudes in parallel so that all cores of the multi-core processors are simultaneously running. The computing time (using the table computer) strongly depends on the required accuracy. For the experimental stack of 4 pancake coils with 24 turns each (96 turns in total) the computing time for 50 elements in the tapes and 40 time steps per cycle is 75 min for the whole AC loss curve with 8 amplitudes, corresponding to less than 10 minutes on average per amplitude. The estimated error by comparing to the results for 100 elements per tape and 320 time steps per cylce is below 4 \%. \subsection{Voltage and scalar potential} \label{s.vphi} Once the current density is known, we can evaluate the voltage drop along each turn in the whole coil as follows. The voltage drop in any situation, including varying magnetic fields, is the difference of the electrostatic potential at the wire ends. Thus, the voltage drop at a certain turn $i$ is $v_i=-2\pi \partial_\varphi\phi_i$, where the sign of the voltage is taken in such a way that the voltage decreases when moving in the direction of the current flow and the current is defined positive if it circulates anti-clockwise. Thus, the total voltage drop in the coil is \begin{equation} v=\sum_{i=1}^{n_t}v_i=-2\pi\sum_{i=1}^{n_t} \partial_\varphi\phi_i. \end{equation} Next, we obtain $\partial_\varphi\phi_i$ from the current density from (\ref{EAphi}), \begin{equation} \partial_\varphi\phi=-r\left [{ E(J)+{\dot A}_J+{\dot A}_a }\right ]. \end{equation} This defines a $\partial_\varphi\phi$ for each position in the tape cross-section. Since $\partial_\varphi\phi$ is actually constant in each turn, it is sufficient to take any arbitrary point in a turn cross-section. However, one may take an average across the turn cross-section in order to minimize the effect of any numerical errors \begin{equation} \partial_\varphi\phi_i=-\frac{1}{S_i}\int_{S_i}{\rm d}s\ r\left [{ E(J)+{\dot A}_J+{\dot A}_a }\right ]. \end{equation} The applied vector potential $A_a$ is typically an analytical function, and thence its time derivative can be calculated straightforwardly. We numerically evaluate ${\dot A}_J$ at a certain time $t_0$, from $A_J$ at $t_0$ and the previous and following time layers, $t_0-\Delta t'$ and $t_0+\Delta t$ respectively, as follows \begin{equation} {\dot A}_J(t_0)\approx \frac{1}{2}\frac{A_J(t_0+\Delta t)-A_J(t_0)}{\Delta t}+\frac{1}{2}\frac{A_J(t_0)-A_J(t_0-\Delta t')}{\Delta t'}. \end{equation} \section{Experimental method} \label{s.expmeth} This section outlines the experimental method for the measurement of the critical current density, the flux creep exponent and the AC loss. \subsection{Critial current density and flux-creep exponent} \label{s.Jcn} In this article, we use a ReBCO coated conductor tape from SuperPower \cite{SuperPower} for all experiments. This tape is 4 mm wide, with a total of 40 $\mu$m copper stabilizer layers, a 1 $\mu$m thick superconducting layer, and a self-field critical current at 77 K of 128 A. We measured the dependence of the critical current density $J_c$ on the magnetic field magnitude $|{\bf B}|\equiv B$ and its orientation $\theta$ (see sketch in figure \ref{f.nBth}) at 77 K, as detailed in \cite{coatedIc}. In order to extract $J_c$ from measurements of the tape critical current, $I_c$, we corrected the spurious effects of the self-field, following the method in \cite{coatedIc}. The reader can find the $I_c$ measurements and extracted $J_c$ for the tape used in this article in \cite{pardo12SSTb}. For completeness, we include the extracted $J_c(B,\theta)$ relation, being \begin{eqnarray} J_c(B,\theta,J) = [ J_{c,ab}(B,\theta,J)^m+J_{c,c}(B)^m ]^{1/m} \label{Jcall} \end{eqnarray} with \begin{eqnarray} J_{c,ab}(B,\theta,J) = \frac{J_{0,ab}}{\left[{1+\frac{Bf(\theta,J)}{B_{0,ab}}}\right]^{\beta_{ab}}}, \label{Jcab} \\ J_{c,c}(B) = \frac{J_{0,c}}{\left[{1+\frac{B}{B_{0,c}}}\right]^{\beta_{c}}}, \label{Jc} \\ \end{eqnarray} and \begin{eqnarray} f(\theta,J) = \left \{ \begin{array}{ll} f_{0}(\theta) & {\rm if}\ J\sin\theta > 0 \\ f_{\pi}(\theta) & {\rm otherwise} \end{array} \right . , \label{f} \\ f_{0}(\theta) = \sqrt{u^2\cos^2(\theta+\delta_{0})+\sin^2(\theta+\delta_{0})}, \label{f0} \\ f_{\pi}(\theta) = \sqrt{u^2\cos^2(\theta+\delta_{\pi})+v^2\sin^2(\theta+\delta_{\pi})}, \label{fpi} \end{eqnarray} where the parameters are $m=8$, $J_{0,ab}=2.53\cdot 10^{10}$ A/m$^2$, $J_{0,c}=2.10\cdot 10^{10}$ A/m$^2$, $B_{0,ab}=414$ mT, $B_{0,c}=90$ mT, $\beta_{ab}=0.934$, $\beta_c=0.8$, $u=5.5$, $v=$1.2, $\delta_0$=-2.5$^{\rm o}$ and $\delta_\pi$=0.5$^{\rm o}$. The critical current density at zero local field is $J_c(B=0)=2.59\cdot 10^{10}$ A/m$^2$. The estimated error of the extracted $J_c(B,\theta)$ is below 5\% \cite{pardo12SSTb}. \begin{figure}[tbp] \begin{center} \includegraphics[width=9cm]{fig3.eps}% \caption{\label{f.nBth} Measured power-law exponent $n$ [see $E(J)$ relation in (\ref{EJ})] directly obtained from tape critical-current measurements. The insert shows a sketch of the angle $\theta$ definition, where the blue rectangle and the black line on top represent the cross-section substrate of the tape and the superconducting layer, respectively.} \end{center} \end{figure} We measured the power-law exponent $n$ in a similar way, although we did not make any self-field correction. The reason is that for low magnetic fields, the $n$ exponent is high ($n$ above 30) and for such high $n$ the electro-magnetic response in alternating currents or magnetic fields weakly depends on this parameter. For the same reason, we assume a periodicity of 90 degrees for the angular dependence. Therefore, we enter directly the measured data in the model (with a bi-linear interpolation in $B$ and $\theta$), since an analytical fit is no longer necessary. \subsection{Coils and AC loss measurement} \label{s.expcoils} We constructed four identical pancake coils of 24 turns each with internal and external diameters 60 and 67.8 mm, respectively, as detailed in \cite{pardo12SSTb} (see figure \ref{f.coils}). Afterwards, we pile the pancakes in stacks of 1 to 4 units, with a total height of 4.0, 8.9, 13.1 and 17.6 mm, respectively. Finally, the AC loss was measured by electrical means as follows. The voltage signal is taken from the taps at the terminals. The transport AC current is measured by a Rogowski coil. We also use this voltage (shifted 90 o with respect to the transport current) to compensate the huge inductive component of the measured voltage of the pancake. We set the desired value of this compensation signal by means of a Dewetron DAQP Bridge-B amplifier (sketch in figure \ref{f.setup}). \begin{figure}[tbp] \begin{center} \includegraphics[height=5cm]{fig4a.eps}% \includegraphics[height=5cm]{fig4b.eps}% \caption{\label{f.coils} Left: one of the pancake coils in the AC loss measurements. Right: stack of 4 pancake coils with mechanical support structure.} \end{center} \end{figure} \begin{figure}[tbp] \begin{center} \includegraphics[width=9cm]{fig5.eps}% \end{center} \caption{\label{f.setup} Set-up to measure the AC loss in the constructed coils.} \end{figure} \section{Benchmarking and comparison with experiments} \label{s.valid} This section tests the numerical method by comparing to analytical formulas for thin strips and experiments on stacks of pancake coils, finding a very good agreement for all cases. \subsection{Single strip} \begin{figure}[tbp] \begin{center} \includegraphics[width=11cm]{fig6.eps}% \caption{\label{f.Kstrip} The sheet current density $K$ from the numerical model agrees with Norris' thin film formula \cite{norris70JPD}. These results are normalized to the critical sheet current density, $K_c=J_cd$, and the horizontal position $x$ is divided to the tape width $w$. The numerical calculations use a power-law exponent $n=1000$ to describe the critical state model.}\label{f.KNorris} \end{center} \end{figure} In this section, we check the numerical model by comparing the results to analytical formulas for thin strips. In order to compare to our method assuming cylindrical symmetry, we take a single-turn coil with radius much larger than the tape width. In particular, we take an inner radius of 1 m, tape width 4 mm (in the $z$ direction) and thickness 1$\mu$m (in the radial direction). First, we compare the sheet current density $K$ (current density integrated over the tape thickness) to Norris' formulas for the critical-state model \cite{norris70JPD} \begin{equation} K(x)= \begin{cases} \frac{2K_c}{\pi}\arctan\sqrt{\frac{(w/2)^2-b^2}{b^2-x^2}} & $for $|x|<b \\ K_c & $for $b<|z|<w/2 \end{cases} \end{equation} with $b=(w/2)\sqrt{1-(I/I_m)^2}$ and $K_c=J_cd$, where $d$ is the strip thickness. The sheet current density from the numerical model for a power-law exponent $n=1000$ coincides to the analytical results (see figure \ref{f.KNorris}). Since the current density agrees with the analytical result, it is not necessary to also compare the AC loss. Actually, checking the current density is more strict than the AC loss, since the current density that produces a given AC loss is not unique. The results in figure \ref{f.KNorris} were computed with 500 elements and a tolerance for $J$ of 0.002 of \% $J_c$. The calculations are for a frequency of 50 Hz, although the results are virtually independent on this parameter. \begin{figure}[tbp] \begin{center} \includegraphics[width=9cm]{fig7.eps}% \caption{The AC loss for a single strip with constant $J_c$ (solid lines with symbols) approaches to the DC limit at high current amplitudes (dash lines). The vertical axis is the dimensionless loss factor $2\pi Q_l/(\mu_0I_m^2)$, where $I_m$ is the current amplitude and $Q_l$ is the loss per cycle and unit length. The results are for a power-law $E(J)$ relation with different $n$ exponents 5,10,20,40,80,200 (in the arrow direction) and frequency 100 Hz. Symbols (and colors) distinguish lines with different $n$.}\label{f.Gnstrip} \end{center} \end{figure} For low power-law exponents, there do not exist analytical formulas for the current density or the AC loss. Nevertheless, the AC loss should approach to the DC limit at high current amplitudes. This DC loss per cycle and unit tape length is \begin{equation} q_{DC}=\frac{S}{T}\oint {\rm d} t E(J_{DC})J_{DC}, \end{equation} where $S$ is the tape cross-section area and the DC current density is $J_{DC}=|I|/S$. For a sinusoidal excitation $I=I_m\sin(\omega t)$ and a power-law $E(J)$ relation, this DC loss becomes \begin{equation} q_{DC}=c(n)E_c \left ( \frac{I_m}{I_c}\right ) ^n I_m \end{equation} with \begin{equation} c(n)\equiv\frac{2}{\pi}\int_0^{\pi/2} {\rm d} \theta (\sin\theta)^{n+1}. \end{equation} For an integer $n$, this function can be evaluated analytically as \begin{equation} c(n) = \begin{cases} \frac{2}{\pi} \frac{[(n/2)!]^2 2^{n}}{(n+1)!} & $if $n$ is even$\\ \frac{(n+1)!}{\{[(n+1)/2]!\}^2 2^{n+1}} & $if $n$ is odd$. \end{cases} \end{equation} In case that $n$ is non-integer $c(n)$ is calculated numerically. Using the DC loss from the equations above, we found that the computed AC loss for a thin strip approaches to the DC limit for high current amplitudes, which supports the validity of the numerical model (see figure \ref{f.Gnstrip}). In that figure, we plot the loss factor $\Gamma=2\pi q/(\mu_0I_m^2)$, where $q$ is the loss per cycle and tape length, in order to emphasize the differences between curves. Figure \ref{f.Gnstrip} also shows that for moderate and high $I_m$ ($I_m$ above around $0.3I_c$) the AC loss increases with increasing $n$, while for low current amplitudes the curves follow the opposite trend. The reasons are the following. For high current amplitudes, the AC loss is mostly originated in the region with $J> J_c$; and thence the same $J$ creates lower $E$ for lower $n$, resulting in a decrease of AC loss with decreasing $n$. On the contrary, for low current amplitudes the contribution to the AC loss from the non-critical region ($J<J_c$) becomes important for low $n$. In that case, $E$ decreases with $n$ for a fixed $J$, and thus the AC loss decreases with $n$ until it saturates for high $n$. A similar behaviour has also been observed for round wires \cite{chen05APL,stenvall10SSTa} and multi-filamentary Bi2223 tapes \cite{stavrevthesis}. The $n$ dependence for a fixed current amplitude have also been studied in \cite{sirois08IES}. \subsection{Comparison with experiments: stack of pancake coils} \label{s.compexp} In the following section, we compare the AC loss calculations with measurements for the experimental stacks of pancake coils (see section \ref{s.expcoils} and \cite{pardo12SSTb}), showing a good agreement. \begin{figure}[tbp] \begin{center} \includegraphics[width=9cm]{fig8.eps}% \caption{\label{f.Qmeas1x24} Computed AC loss with the Critical State Model (CSM) and power-law $E(J)$ relation with different $n$ exponents compared to the measured one for a single pancake coil (frequency 36 Hz). The curve ``$n$ from experiment" corresponds to the data from figure \ref{f.nBth}. A smooth $E(J)$ relation is necessary to describe the over-critical situation. } \end{center} \end{figure} The AC loss for a single pancake coil from (figure \ref{f.Qmeas1x24}) reveals several features. First, the numerical calculations with $n=200$ coincide with those from \cite{pardo12SSTb} for $n=\infty$ (or the critical state model), supporting again the validity of the method presented here. Second, using a smooth power-law $E(J)$ allows predicting the behaviour for transport currents beyond the critical one, which is not possible for the critical-state model. However, the description presented in this article only allows to calculate situations close beyond the critical current. The limitations are due to the fact that for a high enough current, there will be significant current sharing with the stabilization layers and, in addition, the high dissipation will increase the superconductor temperature, eventually experiencing electro-thermal quench behaviour. The agreement of the model with experiments up to relatively high currents (130\% of $I_c$) can be explained by the relatively low $n$ exponent in the $I_c$-limiting turn ($n\approx 18$), reducing the heating and current sharing effects. Third, using the experimental $n(B,\theta)$ relation of figure \ref{f.nBth} provides practically the same loss results than for the minimum $n$ exponent in the measured range, 18.3. The reason is that the AC loss changes little with small changes of $n$ (see curves in figure \ref{f.Gnstrip} for $n=20$ and 40). In addition, the difference in AC loss becomes larger at high $I_m$, in consistence with the behaviour for a thin strip in figure \ref{f.Gnstrip}. Finally, the computed AC loss for the coil agrees with the measurements for all current amplitudes (see figure \ref{f.Qmeas1x24}). The small discrepancies at the highest amplitudes may be due to experimental error in $J_c$ and $n$, the extraction of $J_c$ from measurements, the assumption that $n$ follows a 90 degrees periodicity, and possible partial current flow in the copper stabilization. Next, we discuss the AC loss for the stacks from 1 to 4 pancake coils. Figure \ref{f.Qmeas4x24}a presents the calculations for the critical-state model obtained in \cite{pardo12SSTb}, while figure \ref{f.Qmeas4x24}b is for the results of the model in this work with the measured power-law exponent from figure \ref{f.nBth}. For all sets of pancakes, the calculations with smooth $E(J)$ relation agree better with the experiments than for the critical-state model. The agreement is perfect within the measurement error except for the following cases, where there are slight deviations. First, the model over-estimates the AC loss for stacks of 1 and 2 pancakes at very high currents, with the same causes as discussed above for one single pancake. Second, for a very low amplitudes, there is a slight over-estimation of the AC loss, which may be a consequence of avoiding self-field corrections in $n(B,\theta)$. Then, for low magnetic fields, $n$ is actually larger than the one that the model assumes, slightly over-estimating the AC loss. Finally, the computed loss is slightly above the measured one for the whole curve corresponding to 4 pancakes. This could be caused by non-uniformity in the tape length, so that one of the pancakes exposed to the highest AC loss (top and bottom ones) are made of a tape with slightly larger $J_c$. Additionally, there may also be slight errors in the $J_c$ extraction process. \begin{figure}[tbp] \begin{center} \includegraphics[width=9cm]{fig9a.eps} \\% \includegraphics[width=9cm]{fig9b.eps}% \caption{\label{f.Qmeas4x24} Comparison between measurements and calculations for (a) the critical state model and (b) power-law $E(J)$ relation for the experimental stacks of pancake coils consisting of 1,2,3 and 4 pancakes at 36 Hz frequency. Although the critical-state model agrees with the measurements for mid and low currents, the smooth $E(J)$ relation provides good agreement for all amplitudes, also above the critical current.} \end{center} \end{figure} \section{Magnet-size coils} \label{s.magnet} In this section, we apply our numerical model in order to predict several features of an example of magnet-size coil consisting on a stack of pancake coils, such as those in solenoidal SMES \cite{nomura10IES,sander11SST,zhangH13IES,saichi14IES}, high-field magnets \cite{weijers14IES,awaji14IES} or other solenoidal magnets \cite{maeda14IES,daibo13IES,yoonS14IES}. As a generic example of any of these applications, we study a stack of pancake coils consisting on 20 pancakes with 200 turns per pancake (see table \ref{t.magcoil} and figure \ref{f.magsection}). These calculations serve not only to illustrate the model application but also discuss several features for coated conductor coils with many turns. This section presents a purely modelling analysis, confidently based on the comparison of modelling and experimental results from the previous section. In the following, we present the numerical parameters used in the study (section \ref{s.parnum}), the assumed $J_c(B,\theta)$ relation (section \ref{s.JcBmagnet}), the relaxation effects in $J$ and the generated magnetic field (section \ref{s.rel}), the AC loss (section \ref{s.ACmag}), and the effect of magnetization currents in the magnetic field (section \ref{s.Bqual}). \begin{figure}[tbp] \begin{center} \includegraphics[width=9cm]{fig10.eps}% \caption{\label{f.magsection} The sketch shows the cross-section of the studied coil, where each rectangle represents a pancake coil. Further details in table \ref{t.magcoil}.} \end{center} \end{figure} \begin{table} \begin{center} \footnotesize\rm \begin{tabular}{ll} \hline Number of pancakes & 20 \\ Number of turns per pancake & 200 \\ Inner radius & 29.5 mm \\ Outer radius & 67.2 mm \\ Total heigh & 88.0 mm \\ Tape width & 4.0 mm \\ \hline \end{tabular} \end{center} \caption{Geometrical parameters of the modelled example of magnet-size coil.}\label{t.magcoil} \end{table} \subsection{Numerical parameters} \label{s.parnum} In order to simplify the computations, we take the continuous approximation, that is we approximate each pancake coil as a continuous object with the same engineering current density as the original one \cite{prigozhin11SST,neighbour,zermeno13JAP}. As shown in \cite{neighbour}, this approximation introduces negligible errors, providing a slight under-estimation in the AC loss at very low current amplitudes. With this approximation, we divide the coil radial thickness into 20 equivalent turns of identical cross-section and no separation between them, which transport 10 times the current of one turn in the original 200-turn pancake coil. We also use 50 elements across the tape width and a tolerance of $J$ between 0.008 and 0.002 \% of $J_c(B=0)$, being the lowest values for the lowest current amplitudes or for magnetic relaxation calculations. \subsection{Assumed magnetic field dependence on $J_c$ and coil critical current} \label{s.JcBmagnet} \begin{figure}[tbp] \begin{center} \includegraphics[width=9cm]{fig11.eps}% \caption{\label{f.IcBselva} Magnetic field dependence of $J_c$ used in the model for the coil in figure \ref{f.magsection}. Symbols are measurements at 50 K from Selvanickam {\it et al.} in \cite{selvamanickam12SST} and lines are analytical fits from equations (\ref{f.Ji}) and (\ref{f.Jl}).} \end{center} \end{figure} In order to generate magnetic fields of considerable magnitude, operating temperatures well below 77 K are necessary, due to the severe reduction or the critical current above 1 T at this temperature \cite{daibo13IES,yoonS14IES}. Setting a goal of 7 T of generated magnetic field in the bore, our coil requires a critical current of 190 A, which cannot be achieved at 77 K based on current material performance. Therefore, the experimental $J_c$ data for 77 K in section \ref{s.Jcn} is not useful for this case. Instead, we use the data from \cite{selvamanickam12SST} for 50 K. For simplicity, we took the measured data in that article for applied magnetic fields in the perpendicular and parallel directions and fit the magnetic field dependence with a Kim-like function \cite{kim62PRL}. In addition, we simplified the angular dependence as an elliptical function. Therefore, the assumed magnetic field dependence of $J_c$ is \begin{equation} \label{JcBmag} J_c(B,\theta)=\frac{J_{c0}}{1+\frac{Bf(\theta)}{B_0}} \end{equation} with \begin{equation} \label{Jcthmag} f(\theta)=\sqrt{u^2\cos^2\theta+\sin^2\theta}, \end{equation} where $J_{c0}$, $B_0$, $u$ are constant parameters. The dependence from the equation above (actually $J_cd$) fits well to the experimental data from \cite{selvamanickam12SST} for parallel and perpendicular applied magnetic field for $J_{c0}=1.405\times 10^{11}$ A/m$^2$, $B_0=7.47$ T, $u=5.66$, where we assumed a superconducting layer thickness of $d=1.4$ $\mu$m (see figure \ref{f.IcBselva}). Note that taking $J_c$ directly from $I_c$ measurements, as done for equations (\ref{JcBmag}) and (\ref{Jcthmag}), neglects the self-field effects in the $I_c$ measurements. However, the error in the taken $J_c$ is negligible for magnet-size coils because the local magnetic fields are high. Additionally, although the real angular dependence is more complex than the assumed elliptical type, the obtained results with the above $J_c(B,\theta)$ provide the main features for magnet-size coils. As shown in section \ref{s.compexp}, the numerical model can use a more complex angular dependence, if provided for several applied magnetic fields. With these parameters, we calculated the coil critical current as that of the weakest turn (as defined in previous works \cite{coatedIc,pitel13SST}), with a result of $I_{c,{\rm coil}}=194$ A. In order to be sure that this value corresponds to the DC limit, we have computed the critical current by increasing the current in a quarter sinusoidal cycle of $10^{-14}$ Hz. The critical current is determined by evaluating the voltage per unit length in all the turns (see section \ref{s.vphi}) and using a voltage criterion of 1 $\mu$V/cm. In this article, we arbitrarily chose a power-law exponent of 20, although the method allows to calculate any exponent without significant variation in the computing time. \subsection{Relaxation effects} \label{s.rel} Next, we study the relaxation effects after energizing the coil and keeping the current constant for a certain time. In particular, we analyze the case of increasing the current up to 162 A following a quarter sinusoidal cycle of 0.1 Hz (charging curve of 2.5 s) and afterwards keeping the current constant for one hour. The calculations for this case use a time step that increases exponentially. \begin{figure}[tbp] \begin{center} \includegraphics[width=9cm]{fig12.eps}% \caption{\label{f.Ji} Current density in the coil in figure \ref{f.magsection} at the end of the charging curve, consisting on a quarter sinusoidal cycle of 162 A peak and 0.1 Hz frequency. Negative current density evidences magnetization currents.} \end{center} \end{figure} \begin{figure}[tbp] \begin{center} \includegraphics[width=9cm]{fig13.eps}% \caption{\label{f.Jl} Current density in the coil from figure \ref{f.magsection} for 1 hour relaxation after the charging curve (situation at the end of charging curve in figure \ref{f.Ji}). Magnetization currents are strongly suppressed.} \end{center} \end{figure} At the end of the charging curve, the presence of current density with opposite sign to the transport current is evidence of important magnetization currents (figure \ref{f.Ji}). The 4 top and bottom pancakes are saturated with magnetization currents. After one hour of relaxation, the magnetization currents are strongly suppressed, disappearing from the top and bottom pancakes (see figure \ref{f.Jl}) and the current density becomes more uniform in all pancakes. In more detail, the current density at the end of the charging curve (figure \ref{f.Ji}) presents the same qualitative features as for the critical-state model with constant $J_c$ \cite{neighbour}. Apart from the fact that the pancakes closest to the top and bottom are saturated with magnetization currents, there appears a sub-critical zone in the rest of the pancakes, where $J$ is uniform with roughly the same value for all the pancakes and proportional to the net current in the coil. The main additional feature appearing in the calculations of the present work is that in the critical region (region with $J\ge J_c$) the current density decreases from the edge of the sub-critical zone or border between positive and negative $J$ with approaching the top and bottom edge of each pancake. The cause of this effect is the increase of the radial magnetic field, since it vanishes at the sub-critical zone \cite{pardo12SSTb} and becomes minimum at the border between the positive and negative $J$ due to the magnetic field created by the magnetization currents. \begin{figure}[tbp] \begin{center} \includegraphics[width=9cm]{fig14.eps}% \caption{\label{f.Brel} The relaxation of magnetization currents causes that the generated magnetic field at the bore center increases with time after setting the current to a certain constant value (162 A).} \end{center} \end{figure} The relaxation of current density has an important effect on the generated magnetic field at the bore center (figure \ref{f.Brel}). After one hour relaxation, the generated magnetic field increases by around 100 mT on a background magnetic field of approximately 6 T, representing roughly 1.7 \% increase. This increase is relatively high for magnets and may not be suitable for certain applications, such as Nuclear Magnetic Resonance or accelerator magnets. However, coil optimization can reduce the impact of the relaxation effect. For superconducting tapes with higher power-law exponents, this increase in the generated magnetic field will require higher relaxation times but it will still be present. \subsection{AC loss} \label{s.ACmag} This section discusses the AC loss due to alternating transport currents at 0.1 Hz. \begin{figure}[tbp] \begin{center} \includegraphics[width=9cm]{fig15a.eps} \\% \includegraphics[width=9cm]{fig15b.eps}% \caption{\label{f.Q} (a) Calculated AC loss per cycle $Q$ and (b) computed effective resistance per frequency $R_{\rm eff}=2Q/(fI_m^2)$. For both cases, the AC frequency is 0.1 Hz. The peak in $R_{\rm eff}$ evidences magnetization loss.} \end{center} \end{figure} The AC loss increases with increasing the AC current (see figure \ref{f.Q}a), presenting the following features. At low amplitudes, the loss curve in logarithmic scale presents a slope of around 3; with growing the AC current, the slope decreases down to roughly 1.7; finally, at very high AC currents the slope sharply increases to a value between 20 and 30. The slope of around 3 at low amplitudes and its following decrease can be explained by Bean's slab model for magnetization loss \cite{bean64RMP,goldfarbSpin,clemSpin,reviewac}, since a pancake coil with many turns under a radial applied magnetic field roughly behaves as a slab. In that case, the AC loss is proportional to $H_m^3$, where $H_m$ is the applied magnetic field amplitude, until the slab penetrates. Beyond the $H_m$ where this occurs, the AC loss gradually becomes proportional to $H_m$. Since in a pancake coil of our winding, $H_m$ generated by the other coils is proportional to $I_m$, the loss curve as a function of $I_m$ should follow the same dependence. The fact that in our coil the slope does not decrease as much as 1 is caused by, first, the contribution from the transport loss and dynamic magneto-resistance (as seen for a single tape in \cite{HacIacinphase}) and, second, the onset of the non-linear resistive loss for transport currents below the critical value. Actually, the latter contribution is the responsible of the sharp slope rise at very high currents. A slope higher than the $n$ power-law exponent of 20 at the over-critical situation is caused by the decrease of $J_c$ with the increase of the magnetic field when increasing $I_m$. Note that this contribution to the AC loss is not apparent for currents just above the critical values, requiring $I_m\approx 1.17 I_c$ in our case. The reason is that, according to our definition of coil critical current, at the coil $I_c$ only one turn is above the local $I_c$, and thence the resistive loss contribution to the total AC loss is small. The AC loss behaviour is more evident when represented as the quantity $R_{\rm eff}=2Q/I_m^2$ (see figure \ref{f.Q}b). This quantity has the interpretation of an effective resistance per unit frequency, since the power loss in a device of resistance $R$ is $P=\frac{1}{2}RI_m^2$. With this representation, the saturation of the magnetization loss appears as a peak. \subsection{Magnetic field distortion due to magnetization currents} \label{s.Bqual} In this section, we discuss the effect of the generated magnetic field of the magnetization currents and the distortion that they create. For this purpose, we consider alternating currents of 0.1 Hz frequency. \begin{figure}[tbp] \begin{center} \includegraphics[width=9cm]{fig16.eps}% \caption{\label{f.Bcencyc} Magnetic field at the bore center generated by magnetization currents (total magnetic field minus magnetic field assuming uniform current in the tape) for alternating current of 0.1 Hz. The arrows at the curves for the largest current amplitude represent the current sequence starting from the zero-field-cool situation.} \end{center} \end{figure} First, we analyze the magnetic field at the bore center due to magnetization currents, $B_{\rm c,mag}$, in figure \ref{f.Bcencyc}. This contribution to the generated magnetic field presents a hysteresis cycle with a previous initial curve. With increasing the current, the absolute value of $B_{\rm c, mag}$ at the initial curve first increases, presents a peak, and then decreases. The cause of the initial rise is the creation of magnetization currents, while the reason of the decrease at higher currents is both the decrease of $J_c$ due to the higher local magnetic field and the depletion of magnetization currents due to the increase of transport current. It is important to notice that the remanence is relatively high, up to 330 mT, being 4.5 \% of the maximum generated field at the critical current, 7.3 T. \begin{figure}[tbp] \begin{center} \includegraphics[width=9cm]{fig17.eps}% \caption{\label{f.Bcenin} Contribution of the magnetization currents to the generated magnetic field in the bore center, $B_{c,{\rm mag}}$, relative to the generated magnetic field if magnetization currents are not present, $B_{c,{\rm ideal}}$. The plotted results are for the initial curve.} \end{center} \end{figure} An important parameter for magnet technology is the magnetic field distortion. That is, the relative error of the generated magnetic field compared to the design value. Often, the design value is taken as the magnetic field created by the winding, ignoring magnetization currents \cite{amemiya08SST,ahnMC09IES,yanagisawa11IES,amemiya12PhC} ($B_{\rm ideal})$, and therefore the magnetic field distortion is $|B_{\rm c,mag}|/B_{\rm ideal}$. For our example, this magnetic field distortion at the bore center and the initial curve ranges between 0.13 and 0.018, decreasing with the current (figure \ref{f.Bcenin}). These values are very high for NMR and accelerator magnets \cite{amemiya12PhC,zhangM14SST}, although a decrease of this quality factor could be obtained by optimizing the winding geometry. However, magnets with small bores are likely to present low quality factors, since the magnetic field created by magnetization currents increases with decreasing the distance from the winding. Another way to reduce the magnetic field distortion could be by taking magnetization currents into account in the magnet design. \section{Conclusion} \label{s.conc} Summarizing, this article has presented a method to numerically calculate the electromagnetic properties of superconductors described by any ${\bf E}({\bf J})$ relation under slowly varying magnetic fields. For this purpose, we have obtained a variational principle in the ${\bf J}-\phi$ formulation that reduces the problem to the sample volume, avoiding unnecessary elements in the air; and thence speeding up the computations. Although this formulation is valid for any 3-dimensional shape, the results in this article are for coils with cylindrical symmetry. For this case, we have presented the details of the numerical method to find the current density and other electromagnetic quantities by minimizing the functional. Afterwards, we have satisfactorily tested the method by comparing to thin-strip formulas and experiments for stacks of pancake coils. Finally, we have applied the method to calculate the AC loss, relaxation effects and magnetic field quality of a magnet-size coil made of 20 pancakes and 200 turns per pancake. In particular, we have found that our modelling results coincide with the formulas for thin strips. In addition, the AC loss agrees very well with the measurements of a stack of a few pancake coils and, thanks to the smooth $E(J)$ relation, the loss (and other electromagnetic quantities) can also be predicted for over-critical situations. For the magnet-size coil, we have seen that for a power-law exponent of 20, the magnetization currents are substantially suppressed after 1 hour relaxation, appreciably increasing the generated magnetic field. For higher power-law exponents, the same kind of relaxation will occur but with higher relaxation times. Magnetization currents under cyclic input current are also important, decreasing the magnetic field quality. As a consequence, predicting magnetization currents in coated conductors is necessary for magnet design at least for magnets with small bore or strict specifications regarding the quality factor, such as NMR or accelerator magnets. In conclusion, the modelling tool presented in this article satisfies the requirements to predict the electromagnetic behavior of windings from a few turns to magnet-size coils, as well as other multi-tape arrangements. In addition, the presented variational principle is promising for computationally demanding 3-dimensional problems. \section*{Acknowledgements} We acknowledge L. Prigozhin and M. Kapolka for valuable discussions. The authors acknowledge the use of resources provided by the SIVVP project (ERDF, ITMS 26230120002). The research leading to these results has received funding from the European Union Seventh Framework Programme [FP7/2007-2013] (grant NMP-LA-2012-280432) and the Structural Funds of EU (grant ITMS 26240220088)(0.5).
\subsection*{Introduction }$\,$\\ These notes are from a 3-lecture course given by the author at the ICTP in Trieste, Italy, 1st--5th of September 2014, as part of a graduate summer school on ``$L$-functions and modular forms''. The course is meant to serve as an introduction to $l$-adic Galois representations over local fields with ``$l\neq p$'', and, as the school was followed by a computational workshop, has a somewhat computational bent. It is worth mentioning that the course is not about varieties and their \'etale cohomology, but merely about the representation theory. It should also be said that nothing here is meant to be new --- in fact, it's all pretty old by now, except for perhaps a couple tricks illustrated in the exercises, that were inspired by more recent developments. The prerequisites for the course are fairly modest: the theory of local fields and representation theory of finite groups, such as might be covered in a Masters level course, and, for the sake of examples and motivation, the theory of elliptic curves over local fields (Silverman's book is more than sufficient). Some previous exposure to $L$-functions is desirable, as, after all, this was the main topic of the school. The ``Preliminary exercises'' in the beginning are meant to set the right level, as well as serve as a warm-up for the course. \bigskip\bigskip\bigskip \setcounter{tocdepth}{1} \tableofcontents \newpage \section*{Preliminary exercises \bigskip \subsection*{Representation theory} \begin{ex} Write down the irreducible representations of $S_3$ and $D_{10}$ (throughout $S_n$ (resp. $A_n$) will denote the symmetric (resp. alternating) group on $n$ elements, and $D_{2n}$ will denote the dihedral group of order $2n$). \end{ex} \begin{ex} Using the fact that $S_4/V_4\cong S_3$, find the irreducible representations of $S_4$ (or their characters), where $V_4$ denotes the Klein group. \end{ex} \begin{ex} Compute the induced characters of the two irreducible $2$\--dimensional representations of $D_{10}$ to $A_5$. Compute the character table of $A_5$. \end{ex} \bigskip\bigskip\bigskip \subsection*{Local fields} \setcounter{ex}{0} \begin{ex} Determine the degrees, ramification degrees and residue degrees of $$\Q_2(\zeta_3, \sqrt[3]{3})/\Q_2, \quad \quad \Q_3(\zeta_3, \sqrt[3]{3})/\Q_3\quad\mbox{ and }\quad \Q_7(\zeta_3, \sqrt[3]{3})/\Q_7.$$ Here $\zeta_3$ denotes a primitive cube root of 1. \end{ex} \begin{ex} Compute the corresponding Galois groups, inertia groups, wild inertia groups and Frobenius elements. \end{ex} \bigskip\bigskip\bigskip \subsection*{Elliptic curves} \setcounter{ex}{0} \begin{ex} Show that the elliptic curve $E: y^2+y=x^3-x^2$ has split multiplicative reduction at $11$ and good reduction at all other primes. Compute $\#\tilde{E}(\F_2)$, $\#\tilde{E}(\F_5)$ and $\#\tilde{E}(\F_{11})$. \end{ex} \begin{ex} Find an elliptic curve with additive potentially good reduction at $5$, and another with additive potentially multiplicative reduction at $5$. \end{ex} \newpage \section{Artin representations} Throughout these lectures $K$ will denote a field of characteristic zero. We will abuse notation for representations and use the same symbol both for the homomorphism $G\to \GL_d(\C)$ and for the underlying vector space. \begin{minipage}{0.7\linewidth} \begin{dfn}An \emph{Artin representation} $\rho$ over $K$ is a finite dimensional complex representation of $\Gal(\overline{K}/K)$ that factors through a finite quotient by an open subgroup, i.e.\ there is a finite Galois extension $F/K$ such that $\Gal(\overline{K}/F)\subseteq \ker\rho$, so that $\rho$ corresponds to a representation of $\Gal(F/K)$. \end{dfn} \end{minipage} \begin{minipage}{0.3\linewidth} $$\xymatrix{ \overline{K}\ar^{\Gal(\overline{K}/K)}@/^2pc/@{-}[dd]\ar_{\Gal(\overline{K}/F)}@{-}[d]\\ F\ar_{\Gal(F/K)}@{-}[d]\\ K}$$ \end{minipage} \begin{exa} $K=\Q_5$, $F=\Q_5(\zeta_3, \sqrt[3]{5})$, $$\Gal(F/K)\cong S_3 =\langle s,t|s^3=t^2=id, tst=s^{-1}\rangle$$ \begin{itemize} \item Let $\rho$ be the two dimensional irreducible representation of $S_3$: \begin{minipage}{0.8\linewidth} $$\rho(s)=\abcd{-{1}/{2}}{{{-}\sqrt{3}}/{2}}{{\sqrt{3}}/{2}}{-{1}/{2}}\quad \rho(t)=\abcd{1}{0}{0}{{-}1}.$$ \end{minipage} \begin{minipage}{0.2\linewidth} \begin{tikzpicture} \draw (-1,0) -- (1,0); \draw (0,-1) -- (0,1); \draw (-1/3,-1/2) -- (-1/3,1/2)--(2/3,0)--(-1/3,-1/2); \end{tikzpicture} \end{minipage} It gives an irreducible representation of $\Gal(\overline{K}/K)$ by$$ \xymatrix{\Gal(\overline{K}/K)\ar@{>>}[r]& \Gal(F/K)\ar[r]& \GL_2(\C).}$$ \item Let $\epsilon$ be the sign representation of $S_3$: $\epsilon(s)=1$ and $\epsilon(t)={-}1$. This gives a $1$\--dimensional Artin representation of $\Gal(\overline{K}/K)$ which is the same Artin representation as when taking $F=\Q_5(\zeta_3)$ and the non\--trivial $1$\--dimensional representation $\Gal(F/K)\cong C_2\to \GL_1(\C)=\C^\times$. \item The trivial representation $\rho=\mathbf{1}$ gives the same Artin representation over~$K$ for all $F/K$. \end{itemize} \end{exa} \underline{Notation}: $\,$\\ When $K/\Q_p$ is finite and $F/K$ is a Galois extension, we write: \begin{itemize} \item $\pi_K$ for a fixed uniformizer of $K$; \item $\O_K$ for the ring of integers of $K$; \item $v_K$ for the normalized valuation on $K$; \item $\F_K$ for the residue field of $K$ (of characteristic $p$), with $q=\#\F_K$; \item $I_{F/K}$ for the inertia group, i.e.\ the subgroup of $\Gal(F/K)$ that acts trivially on the residue field of $F$: $$I_{F/K}\colon=\left\{g\in \Gal(F/K): g(x)\equiv x \bmod \pi_F \;\forall x\in \O_F\right\},$$ when $F/K$ is finite. \item $\Frob_{F/K}$ for a Frobenius element, i.e.\ any $g\in \Gal(F/K)$ that induces the Frobenius automorphism $x\mapsto x^q$ on the residue field of $F$. \item $\Phi_{F/K}=\Frob_{F/K}^{{-}1}$ for the ``geometric Frobenius". \end{itemize} \begin{dfn} For an Artin representation $\rho$ over a local field $K$ its \emph{local polynomial} is $$P(\rho, T)=\det(1{-}\Phi_{F/K}T|\rho^{I_{F/K}})$$ where $\rho$ factors through $F/K$ and $\rho^{I_{F/K}}$ is the subspace of $\rho$ of $I_{F/K}$\--invariant vectors. \end{dfn} Note that as $I_{F/K}$ is a normal subgroup, $\rho^{I_{F/K}}$ is a subrepresentation of $\rho$, and that $P(\rho, T)$ is essentially the characteristic polynomial of $\Phi_{F/K}$ on this subspace. \bigskip \bigskip \begin{exa} Let $K=\Q_5$ and $F=\Q_5(\zeta_3, \sqrt[3]{5})$, as in the previous example. Here $I_{F/K}\cong C_3=\Gal(F/K(\zeta_3))$ and $\Frob_{F/K}=t$, an element of order $2$.\\ For the trivial representation, $P(\mathbf{1}, T)=\det(1-\triv(t)T|\triv^{I_{F/K}})=\det(1{-}T|\mathbf{1})=1-T$.\\ For the sign representation, $P(\epsilon, T)=\det(1-\epsilon(t)T|\epsilon^{I_{F/K}})=\det(1{+}T|\epsilon)=1+T$.\\ For the 2-dimensional irreducible representation of $S_3$, we find that $P(\rho, T)=1$, since $\rho^{I_{F/K}}=0$. \end{exa} \bigskip \begin{dfn}For an {Artin representation} $\rho$ over a number field $K$, its \emph{Artin L-function} is $$L(\rho, s)=\prod_{\p\subset \O_K {\text{ prime}}}\frac{1}{P_\p(\rho, \nm(\p)^{-s})},$$ where $P_\p(\rho, T)$ is the local polynomial for $\rho$ restricted to $\Gal(\overline{K}_\p/K_\p)$. The Euler product is known to converge for $\Re(s)>1$ to an analytic function. \end{dfn} \bigskip \begin{exa} If $\rho=\mathbf{1}$, then $P_\p(\rho, T)=1-T$ for all $\p$, so we obtain the Dedekind $\zeta$\--function of $K$: $$L(\mathbf{1},s)=\prod_{\p}\frac{1}{ 1-\nm(\p)^{-s}}=\zeta_K(s).$$ Let $K=\Q$ and $\rho$ be the order $2$ character of $\Q(\zeta_3)/\Q$. Then $P_{(3)}(\rho, T)=1$, $P_p(\rho, T)=1\!-\!T$ for $p\equiv 1 \pmod 3$ and $P_p(\rho, T) =1\!+\!T$ for $p\equiv 2 \pmod 3$. So $$L(\rho, s)=\prod_{p\neq 3}\frac{1}{1-\left(\frac{p}{3}\right)p^{-s}}=\sum_{n=1}^\infty\left(\frac{n}{3}\right) n^{-s}= L\left(\left(\frac{\cdot}{3}\right),s\right),$$ the $L$-function of the non-trivial Dirichlet character $\Z/3\Z\to \C^\times$. \end{exa} Fact: Artin $L$\--functions of $1$\--dimensional Artin representations over $\Q$ correspond to Dirichlet $L$\--functions of primitive characters. \subsection*{Basic properties} (i) For $\rho_1$ and $\rho_2$ Artin representations over a local field $K$, we clearly have $P(\rho_1\oplus\rho_2, T)=P(\rho_1, T) P(\rho_2, T)$. (ii) When $F/K$ is a finite extension and $\rho$ is an Artin representation over $F$, then $P_{F}(\rho, T^{f})=P_{K}(\Ind\rho, T)$ where $f$ is the residue degree of $F/K$ and $P_F(\rho, T)$ is the local polynomial for the Artin representation $\rho$ over $F$ and analogously for $P_{K}(\Ind\rho, T)$. (iii) When $K$ is a number field, $L(\rho_1\oplus\rho_2, s)=L(\rho_1, s) L(\rho_2, s)$. If $F/K$ is a finite extension and $\rho$ an Artin representation over $F$, then $L(\rho, s)=L(\Ind\rho, s)$, where the first is an Artin $L$-function over $F$ and the second over $K$. \begin{rmk} Artin $L$\--functions of $1$\--dimensional representations are known to be analytic on $\C$ (except for a pole at $s=1$ for $\rho=\mathbf{1}$). Brauer's induction theorem taken together with (iii) then shows that all Artin $L$\--functions are meromorphic. \begin{conj}[Artin] Let $\rho\neq \mathbf{1}$ be an irreducible Artin representation over a number field. Then its $L$\--function is analytic. \end{conj} \end{rmk} \begin{dfn} For an Artin representation $\rho$ over a local field $K$, its \emph{conductor exponent} $n_{\rho}$ is $$n_{\rho}=n_{\rho,\mbox{\scriptsize tame}}+n_{\rho,\mbox{\scriptsize wild}}$$ with $$n_{\rho,\mbox{\scriptsize tame}}=\dim \rho-\dim \rho^I=\dim \rho/\rho^I,$$ and $$n_{\rho,\mbox{\scriptsize wild}}=\sum_{k=1}^\infty \frac{1}{[I:I_k]} \dim \rho/\rho^{I_k};$$ where $\rho$ factors though $\Gal(F/K)=G$, and $I=I_{F/K}=I_0$ and $$I_k=\{\sigma\in G: \sigma(\alpha) \equiv \alpha\bmod \pi_F^{k+1}\;\;\forall \alpha \in \O_F\}$$ are the higher ramification groups (with the lower numbering). So, in particular, $$I_1={\rm Syl}_p I = \mbox{ wild inertia, }\;\qquad I/I_1=\mbox{ tame inertia (cyclic)}.$$ We say that $\rho$ is \emph{unramified} (resp. \emph{tame}) if $n_\rho=0$ (resp. if $n_{\rho,\mbox{\scriptsize wild}}=0$), equivalently, if $I$ (resp. $I_1$) acts trivially on $\rho$. The \emph{conductor} of $\rho$ is the ideal $\cnd_\rho=(\pi^{n_{\rho}})$. \end{dfn} \begin{rmks} We clearly have: $n_{\rho_1 \oplus\rho_2}=n_{\rho_1}+n_{\rho_2}$ and $\cnd_{\rho_1\oplus\rho_2}=\cnd_{\rho_1}\cnd_{\rho_2}$. One can show that when $\rho$ is irreducible and ramified, $$n_{\rho,\mbox{\scriptsize wild}}=\dim(\rho)\cdot \max\{i: G^i \;\mbox{ acts non\--trivially on } \rho\},$$ where $G^i$ is the $i$\--th ramification group in the upper numbering. \end{rmks} \pagebreak \begin{thm}[Swan's character] Let $\rho$ be an Artin representation of a local field $K$, which factors through $\Gal(F/K)$. Then $$n_{\rho,\mbox{\scriptsize wild}}=\langle \Tr\rho, b\rangle,$$ where \begin{equation*} b(g)= \left\{\begin{array}{ll} 1-v_F(g(\pi_F)-\pi_F) & \mbox{ for } g \in I\setminus \{e\}\\ -\sum_{h\neq e}b(h) & \mbox{ for } g=e, \end{array}\right. \end{equation*} $v_F$ is the normalized valuation of $F$ and $\langle\cdot,\cdot\rangle$ is the representation theoretic inner product for characters of $I_{F/K}$. \end{thm} \begin{thm}[Artin]$n_\rho\in \Z$.\end{thm} \begin{thm}[Conductor\--discrimant formula]$\,$\\ \begin{minipage}{0.2\linewidth} $$ \xymatrix{ F\ar^{H}@{-}[rd] \ar_{G}@{-}[dd] & \;\\ \; & L \ar@{-}[dl] \\ K &\, }$$ \end{minipage} \begin{minipage}{0.8\linewidth} Let $F/K$ be a Galois extension with an intermediate field $L$, and let $\rho$ be a representation of $H=\Gal(F/L)$. Then $$n_{\Ind_H^G \rho}=(\dim \rho)\cdot v_K(\Delta_{L/K})+f_{L/K} \cdot n_\rho.$$ Equivalently, $$\cnd_{\Ind \rho}=\Delta_{L/K}^{\dim \rho}\cdot \nm_{L/K}(\cnd_{\rho}),$$ where $f_{L/K}$ is the residue degree, $\Delta_{L/K}$ is the relative discriminant, and $\nm_{L/K}$ the relative norm. \end{minipage} \end{thm} \begin{exa}$\,$\\ \begin{minipage}{0.4\linewidth} $$ \xymatrix@dr@C=1pc{ F\ar@{-}[r] \ar@{-}[d] &L=\Q_5(\sqrt[3]{5}) \ar@{-}[l] \ar@{-}[d] \\ \Q_5(\zeta_3) \ar@{-}[r] &K=\Q_5 }$$ \end{minipage} \begin{minipage}{0.6\linewidth} $I_{F/K}\cong C_3=\Gal(F/K(\zeta_3))$\\ $I_1=\{1\}$\\ $\,$\\ $n_{\mathbf{1}}=0$\\ $n_\epsilon=0$\\ $n_{\rho,\mbox{\scriptsize tame}}=2-0=2, \quad n_{\rho,\mbox{\scriptsize wild}}=\sum 0=0,\quad n_\rho=2.$ \end{minipage} In particular, by the conductor-discriminant formula, $$1 \cdot \Delta^1_{L/K}=\cnd_{\Ind_{C_2}^{S_3}\mathbf{1}}=\cnd_\rho\cdot \cnd_{\mathbf{1}}=5^2 \;\mbox{up to units,}$$ $$\Delta_{F/K}=\cnd_{\Ind_{1}^{S_3}\mathbf{1}}=\cnd_{\rho\oplus \rho\oplus \epsilon \oplus \mathbf{1}}=5^4 \;\mbox{up to units}.$$ \end{exa} \bigskip \begin{dfn} The \emph{conductor} of an Artin representation $\rho$ over a number field $K$ is $$\cnd_\rho=\prod_{\p \mbox{\scriptsize{\emph{ prime in }}}K}\p^{n_\p(\rho)},$$ where $n_\p(\rho)$ is the conductor exponent of $\rho$ restricted to $\Gal(\overline{K}_\p/K_\p)$.\end{dfn} \bigskip \begin{thm}The Artin $L$-function of $\rho$ satisfies the functional equation: $$\Lambda(\rho, s)=w \cdot A^{\frac{1}{2}-s} \cdot \Lambda(\rho^\ast, 1-s),$$ where $$\Lambda(\rho, s)= L(\rho,s) \cdot \prod_{\nu \mbox{\emph{\scriptsize{ real}}}}\Gamma_\R(s)^{d_{+}(\rho)}\Gamma_\R(s+1)^{d_{-} (\rho)} \prod_{\nu \mbox{\scriptsize{\emph{ complex}}}}\Gamma_\C(s),$$ $\rho^\ast$ denotes the dual representation, $d_{\pm}(\rho)$ is the dimension of the $\pm$\--eigenspace of the image of complex conjugation at~$\nu$, $w\in \C^\ast$ with $|w|=1$ is the \emph{global root number} and \begin{eqnarray*} A&=& \nm(\cnd_\rho)\cdot \sqrt{|\Delta_K|}^{\;\dim \rho},\\ \Gamma_\R(s)&=& \pi^{{-}\frac{s}{2}}\Gamma(s/2),\\ \Gamma_\C(s)&=&(2\pi)^{{-}s}\Gamma(s). \end{eqnarray*} \end{thm} \newpage \section*{Exercises to Lecture 1 \bigskip \setcounter{ex}{0} \begin{ex} For each irreducible Artin representation over $\Q$ that factors through $\Q(\zeta_3, \sqrt[3]{3})$ determine its conductor and the first $5$ coefficients of its Artin L\--series $\sum_{n \geq 1} a_n n^{-s}$. \end{ex} \bigskip\bigskip\begin{ex}$\,$ \\ \begin{minipage}{0.2\linewidth} $$ \xymatrix@dr@C=1pc{ F\ar@{-}[r] \ar@{-}[d] \ar@{-}^{10}[dr] & L \ar@{-}[l] \ar@{-}^{5}[d] \\ K \ar@{-}_ {2}[r] &\Q }$$ \end{minipage} \begin{minipage}{0.8\linewidth} Suppose $F/\Q$ is a Galois extension with Galois group $D_{10}$, the dihedral group of order $10$. Let $K$ and $L$ denote intermediate fields with $[K:\Q]=2$ and $[L:\Q]=5$. Prove the identity $$\zeta_F(s)\zeta_\Q(s)^2=\zeta_L(s)^2\zeta_K(s).$$ \end{minipage} \end{ex} \bigskip\bigskip \begin{ex} The polynomial $f(x)=x^5+2x^4-3x^3+1$ has Galois group $D_{10}$. Its complex roots are $$\alpha_1=-3.01\dots, \qquad \alpha_2=-0.35 \ldots-0.53\dots i, \qquad \alpha_3=0.85\ldots-0.31\dots i,$$ $$\alpha_4=\overline{\alpha_3}, \qquad \alpha_5=\overline{\alpha_2},$$ and the Galois group contains the $5$\--cycle $(\alpha_1 \alpha_2\alpha_3\alpha_4\alpha_5)$. Prove that the Frobenius element of a prime above $2$ in the splitting field of $f(x)$ has order $5$ and determine its conjugacy class in the Galois group.\\ Hint: look at $$\alpha_1\alpha_2+\alpha_2\alpha_3+\alpha_3\alpha_4+\alpha_4\alpha_5+\alpha_5\alpha_1,$$ $$\alpha_1\alpha_3+\alpha_3\alpha_5+\alpha_5\alpha_2+\alpha_2\alpha_4+\alpha_4\alpha_1,$$and $$\beta_1\beta_2+\beta_2\beta_3+\beta_3\beta_4+\beta_4\beta_5+\beta_5\beta_1,$$ where $\beta_1$ is a root of $f(x)$ in $\overline{\F}_2$ and $\beta_i=\beta_1^{2^{i-1}}$. \end{ex} \newpage \section{$l$-adic representations} \begin{dfn}A continuous $l$\--adic representation over $K$ is a continuous homomorphism $\Gal(\overline{K}/K)\to \GL_d(\mathcal{F})$ for some finite extension $\mathcal{F}/\Q_l$. \end{dfn} \bigskip \begin{rmk} An $l$\--adic representation is continuous if an only if for all $n$ there exists a finite Galois extension $F_n/K$ such that $\Gal(\overline{K}/F_n)\to \Id\; \bmod \;l^n$, i.e.\ such that $\rho \bmod \;l^n$ factors through a finite extension $F_n/K$ (except that the image of $\rho$ may have denominators, so $\rho\bmod l^n$ may not actually be well-defined). An Artin representation $\rho$ that factors through $F/K$ can be realized over $\overline{\Q}$, and hence also over a finite extension $\mathcal{F}/\Q_l$. It maps $\Gal(\overline{K}/F)$ to $\Id$, so we can take $F_n=F$ for all $n$ to see that it is continuous. \end{rmk} \bigskip \begin{exa}[Cyclotomic character] Let $\zeta_{l^n}$ be primitive $l^n$\--roots of unity in $\overline{K}$ with $(\zeta_{l^n})^l=\zeta_{l^{n{-}1}}$. For $g\in \Gal(\overline{K}/K)$ define a sequence of integers $0\le a_i <l$~by \begin{eqnarray*} g(\zeta_l)&=&\zeta_l^{a_1},\\ g(\zeta_{l^2})&=&\zeta_{l^2}^{a_1+l\, a_2},\\ &\dots &\\ g(\zeta_{l^n})&=&\zeta_{l^n}^{a_1+l\, a_2+\cdots+l^{n-1}a_n}. \end{eqnarray*} We then define the {\em $l$-adic cyclotomic character} $\chi_{cyc}$ by $$ \chi_{cyc}(g)=a_1+l\, a_2+\cdots+l^{n-1}a_n+\cdots\in\Z_l. $$ Note that the value $\chi_{cyc}(g)\bmod l^n$ simply says what $g$ does to the $l^n$\--th roots of $1$. It is easy to check that the $l$\--adic cyclotomic character is multiplicative, and hence gives a 1-dimensional representation, $$\chi_{cyc}: \Gal(\overline{K}/K)\to \Z_l^\ast\subset\GL_1(\Q_l).$$ Taking $F_n=K(\zeta_{l^n})$, we have that $\Gal(\overline{K}/F_n)\to \Id \bmod\, l^n$, so $\chi_{cyc}$ is continuous. \end{exa} \bigskip \begin{exa}[Tate module] Let $E/K$ be an elliptic curve and let $P_n, Q_n$ be a basis for $E[l^n]$ with $lP_n=P_{n-1}$ and $lQ_n=Q_{n-1}$. For $g\in \Gal(\overline{K}/K)$, we define $0\le a_i, b_i, c_i, d_i<l$ by $g(P_1)= a_1 P_1 + b_1 Q_1$ and $g(Q_1)=c_1 P_1+ d_1 Q_1$, and generally $$ \begin{array}{cccccccccc} g P_n &=& (a_1 + \ldots + a_nl^{n-1}) P_n &+ &(b_1 + \ldots + b_nl^{n-1}) Q_n, \cr g Q_n &=& (c_1+\ldots+c_nl^{n-1}) P_n &+ & (d_1+\ldots +d_nl^{n-1}) Q_n. \end{array} $$ Then $$ \rho(g)=\abc{a_1+\ldots+l^{n-1}a_n+\ldots\quad}{c_1+\cdots+l^{n-1}c_n+\cdots}{b_1+\ldots+l^{n-1}b_n+\ldots\quad}{d_1+\ldots+l^{n-1}d_n+\ldots} \in \GL_2(\Z_l)\subseteq \GL_2(\Q_l) $$ is the representation on the $l$-adic Tate module of $E$. Here we can take $F_n=K(E[l^n])$ to see that $\rho$ is continuous, and the value of $\rho(g) \bmod l^n$ says what $g$ does to $E[l^n]$. \end{exa} \newpage \begin{rmks}$\,$ \begin{enumerate}[(i)] \item Let $K/\Q_p$ be a finite extension and $\mathcal{F}/\Q_l$ with $l\neq p$. Then $\Frob_{\overline{K}/K}(\zeta_{l^n})=\zeta_{l^n}^q$ and $I_{\overline{K}/K}$ acts trivially on $\zeta_{l^n}$ for all $n$. In other words $$\chi_{cyc}(I_{\overline{K}/K})=1 \qquad {\text{and}} \qquad \chi_{cyc}(\Frob_{\overline{K}/K})=\#\F_K=q.$$ \item If $\rho$ is the representation on the Tate module of an elliptic curve, then $\det \rho=\chi_{cyc}$, i.e.\ the determinant of the matrix $\rho(g)$ simply depends on the action of $g$ on $\zeta_{l^n}$ for all $n$. This is because the Weil pairing $E[l^n]\times E[l^n]\to\mu_{l^n}$ is alternating and Galois equivariant (and is, effectively, the determinant map). \item For elliptic curves, when $K/\Q_p$ is a finite extension and $l\neq p$, one usually uses the dual representation $$\rho_E=\rho^\ast,$$ i.e.\ $\rho_E(g)=(\rho(g)^{-1})^t$; this is the same as the representation on $\coh^1_{\mbox{\scriptsize{\emph{\'et}}}}(E, \Q_\ell)$. \end{enumerate} \end{rmks}\bigskip \begin{dfn} Let $K/\Q_p$ be a finite extension and $\rho:\Gal(\overline{K}/K)\to \GL_d(\mathcal{F})$ a continuous $l$\--adic representation with $l\neq p$. The \emph{local polynomial} of $\rho$ is $$P(\rho, T)=\det(1{-}\Phi_{\overline{K}/K} T|\rho^{I_{\overline{K}/K}}),$$ and its {\em conductor exponent} is $n_{\rho}=n_{\rho,\mbox{\scriptsize tame}}+n_{\rho,\mbox{\scriptsize wild}}$ with \begin{eqnarray*} n_{\rho,\mbox{\scriptsize tame}}&=&\dim \rho/\rho^{I_{\overline{K}/K} },\\ n_{\rho,\mbox{\scriptsize wild}}&=&\sum_{k\geq 1} \frac{1}{[I_{F/K}:I_{F/K, k}]} \dim \rho/\rho^{I_{F/K, k}}, \end{eqnarray*} where $F/K$ is a finite extension, large enough so that the action of the wild inertia group factors through $F/K$. (This exists: one can take $F=F_1$ so that the image of $\Gal(\overline{K}/F)$ lies in $\abcd{1{+}l\mathcal{O_F}}{l\mathcal{O_F}}{l\mathcal{O_F}}{1{+}l\mathcal{O_F}}$. The image of wild inertia is trivial since it is a (pro) $p$\--group and this matrix group a (pro) $l$\--group). The {\em conductor} of the representation $\rho$ is $\cnd_\rho=(\pi_K)^{n_{\rho}}$. \end{dfn} \begin{dfn}For an elliptic curve $E$ over a number field $K$, its \emph{L-function} is defined by the Euler product $$L(E, s)=\prod_{\p \mbox{ \scriptsize{\emph{prime}}}}\frac{1}{P_\p(\rho_E, \nm(\p)^{-s})},$$ where $P_\p(\rho_E, T)$ is the local polynomial for $\rho_E:\Gal(\overline{K}_\p/K_\p)\to \GL_2(\Q_l)$ for any $l$ not divisible by $\p$. It converges for $\Re(s)\gg 1$. Its conductor is $\cnd_E=\prod_\p \p^{n_{\p,\rho_E}}$ and it conjecturally satisfies the functional equation: $$\Lambda(\rho, s)=w \cdot A^{1-s}\cdot \Lambda(\rho, 2-s),$$ where \begin{eqnarray*} \Lambda(\rho, s)&=&L(E,s) \Gamma_\C(s)^{[K:\Q]},\\ \Gamma_\C(s)&=&(2\pi)^{{-}s}\Gamma(s),\\ A&=& \nm_{K/\Q}(\cnd_E)\cdot |\Delta_K|^2,\\ w&=&\pm 1 \mbox{ is the \emph{global root number} of }E/K . \end{eqnarray*} \end{dfn} \begin{rmk}This construction of $L$\--functions applies more generally to ``compatible systems of $l$-adic representations " (e.g.\ from Artin representations, abelian varieties, $\coh^i_{\mbox{\scriptsize{\emph{\'et}}}}$ of varieties, motives). The exact formula for the $\Gamma$\--factors and the root number $w$ is explicitly given using the associated Hodge structure and the theory of local $\epsilon$\--factors of $l$-adic representations. \end{rmk} \section{The $l$-adic representation of an elliptic curve} Recall that to an elliptic curve $E/K$ we have associated an $l$-adic representation $\rho_E$. We now relate its properties to the arithmetic of $E$ in the case when $K$ is a local field. \begin{thm} Let $K/\Q_p$ be a finite extension, $E/K$ an elliptic curve, and let \hbox{$\rho_E:\Gal(\overline{K}/K)\to \GL_2(\Q_l)$} the $l$-adic representation on $\coh^1_{\mbox{\scriptsize{\emph{\'et}}}}(E, \Q_\ell)$ with $l\neq p$. Then \begin{enumerate}[1.] \item $E$ has good reduction if and only if $\rho_E$ is unramified (the N\'eron\--Ogg\--Shafarevich criterion), \item $\det\rho_E=\chi_{cyc}^{-1}$ (by the Weil pairing), \item $P(\rho_E, {1}/{q})={\#\tilde{E}(\F_K)}/{q}$, where $q=\#\F_K$. \end{enumerate} \end{thm} \bigskip \begin{rmks} \begin{enumerate}[(i)] \item By $(1)$ $E$ has potentially good reduction if and only if $I_{\overline{K}/K}$ acts through a finite quotient. \item By $(1), (2)$ and $(3)$, if $E/K$ has good reduction then $$P(\rho_E, T)=1-aT+qT^2\quad\mbox{with}\quad a=1+q-\#\tilde{E}(\F_K).$$ \item By $(1)$ and $(3)$: \begin{eqnarray*} E \mbox{ has additive reduction } &\Rightarrow & P(\rho_E, T)=1,\\ E \mbox{ has split multiplicative reduction } &\Rightarrow & P(\rho_E, T)=1-T,\\ E \mbox{ has non\--split multiplicative reduction } &\Rightarrow & P(\rho_E, T)=1+T. \end{eqnarray*} \end{enumerate} \end{rmks} \newpage \begin{exa}$\,$ \begin{minipage}{0.6\linewidth} Let $E/\Q_5$ be the elliptic curve $y^2=x^3+5^2$. It has additive reduction ($\Delta_E= \mbox{unit} \cdot 5^4$). It has good reduction over $L=\Q_5(\sqrt[3]{5})$: over $L$ we have the model $E': y^2=x^3+1$ with $\tilde{E}'(\F_5)=\{\O, (0,\pm 1),(2, \pm2),({-}1,0) \}$. The action of $\Gal(\overline{\Q}_5/\Q_5)$ factors through $\Gal(L^{nr}/\Q_5)$ which is generated by $g\in \Gal(L^{nr}/\Q^{nr}_5)$ of order $3$ (inertia group) and $\Phi_L=\Phi_{L^{nr}/L}$, the geometric Frobenius over $L$. They satisfy the relation $$\Phi_L \cdot g \cdot \Phi_L^{-1}=g^{-1}.$$ \end{minipage} \begin{minipage}{0.4\linewidth} $$ \xymatrix@dr@C=1pc{ \overline{\Q}_5 \ar@{-}[r] & L^{nr} \ar^{C_3=\langle g \rangle}@{-}[r] \ar@{-}[d] & \Q^{nr}_5\ar@{-}[dd] \\ \; & L(\zeta_3)\ar@{-}[d] \ar^{S_3}@{.}[dr] &\;\\ \; & L\ar^3 @{-}[r]\ar^{\langle \Phi_L \rangle}@/^2pc/@{-}[uu] &\Q_5}$$ \end{minipage} In particular $\rho_E(g)$ has eigenvalues $\zeta_3, \zeta_3^{-1}$ (it is nontrivial as there is bad reduction, so it has order $3$, and it has determinant $1$ by the Weil pairing) and $\rho_E(\Phi_L)$ has eigenvalues $\pm \sqrt{{-}5}$ (as $\#\tilde{E}'(\F_5)=6$ so $P(\rho_E, T)=1+5T^2$ over $L$); and $$\rho_E(\Phi_L) \cdot \rho_E( g) \cdot \rho_E(\Phi_L)^{-1}=\rho_E(g)^{-1}.$$ A little algebra shows that, with a suitable choice of $\overline{\Q}_l$\--basis, the representation $\rho_E: \Gal(L^{nr}/K)\to \GL_2(\overline{\Q}_l)$ is given by $$\rho_E(g)=\abcd{\zeta_3}{0}{0}{\zeta_3^{-1}} \quad\quad\quad \mbox{ and } \quad\quad\quad\rho_E(\Phi_L)=\abcd{0}{\sqrt{{-}5}}{\sqrt{{-}5}}{0}.$$ \end{exa} \newpage \section*{Exercises to Lecture 2 \bigskip \setcounter{ex}{0} \begin{ex} Let $E/\Q_7$ be the elliptic curve $y^2+y=x^3-x^2$. Let $K/\Q_7$ be some horrible extension with residue degree $11$ and ramification degree $76$. Find $\#\tilde{E}(\F_K)$, where $\F_K$ is the residue field of $K$. \end{ex} \bigskip \bigskip \begin{ex} Let $K/\Q_p$ be a finite extension and $E/K$ an elliptic curve. Prove that if $p\neq 2,3$ then the exponent of the conductor of $E$ is at most $2$. Hint: pick $\ell\neq p$ that forces the image of the wild inertia group in $\GL_2(\Z/\ell^n\Z)$ to be trivial for all $n$. \end{ex} \bigskip \bigskip \begin{ex} Let $E/\Q_7$ be the elliptic curve $y^2=x^3+7^2$. Identify its $\ell$\--adic representation $\rho_E:\Gal(\overline{\Q}_7/\Q_7)\to \GL_2(\overline{\Q}_\ell)$ on $\coh^1_{\mbox{\scriptsize{\emph{\'et}}}}(E, \Q_\ell)\otimes_{\Q_\ell}\overline{\Q}_\ell$ for $l\neq 7$. Hint: compute $\#\tilde{E}(\F_K)$ over two different cubic ramified extensions $K/\Q_7$. \end{ex} \newpage \section{Classification of $l$-adic representations} Throughout this section $K$ is a finite extension of $\Q_p$ and $l\neq p$ a prime. \begin{exa} $\,$\\ \begin{minipage}{0.6\linewidth} In the last lecture we had $E/\Q_5$ with $l$\--adic representation $\rho_E$ given by $\rho_E(g)=\abcd{\zeta_3}{0}{0}{\zeta_3^{-1}}$ and $\rho_E(\Phi_{L})=\abcd{0}{\sqrt{{-}5}}{\sqrt{{-}5}}{0}$, where $\langle g \rangle =I\cong C_3$ with $g( \sqrt[3]{5})= \zeta_3 \sqrt[3]{5}$, and $\Phi_{L}=\Phi_{L^{nr}/L}$ is a topological generator of $\Gal(L^{nr}/L)$. The representation $\rho_E$ can clearly be written as $$\rho_E=\rho\otimes \psi,$$ where $\rho$ is the $2$\--dimensional representation of $\Gal(\Q_5(\zeta_3, \sqrt[3]{5})/\Q_5)\cong S_3$: $$\rho(g)=\abcd{\zeta_3}{0}{0}{\zeta_3^{-1}}, \quad \rho_E(\Phi_{L})=\abcd{0}{1}{1}{0},$$ \end{minipage} \begin{minipage}{0.4\linewidth} \xymatrix@dr@C=1pc{ \overline{\Q}_5 \ar@{-}[rr] & \, & \Q^{nr}_5(\sqrt[3]{5}) \ar@{-}[r]^(0.65){C_3=\langle g \rangle} \ar@{-}[d] & \Q^{nr}_5\ar@{-}[d] \\ \; &\; & L(\zeta_3) \ar@{-}[d] \ar@{.}[dr]^(0.35){S_3}\ar@{-}[r] & \Q_5(\zeta_3)\ar@{-}[d] \\ \; & \; &L{=}\Q_5(\sqrt[3]{5})\ar^{\langle \Phi_{L} \rangle}@/^2pc/@{-}[uu] \ar @{-}[r] &\Q_5} \end{minipage} and $\psi$ satisfies $\psi(I)=1$ and $\psi(\Phi_{\overline{\Q}_5/\Q_5})={\sqrt{{-}5}}$. This decomposition works as $\rho_E(\Phi_L)^2=\abcd{-5}{0}{0}{-5}$ is a scalar matrix. Indeed, $\Phi_L^2$ is central in $\Gal( \Q^{nr}_5(\sqrt[3]{5})/\Q^{nr}_5)$ and $\rho_E$ is irreducible, so by Schur's lemma $\rho_E(\Phi_L)^2$ must be scalar. This trick works in general for irreducible representations, and lets one prove the following classification: \end{exa} \bigskip \begin{thm} Every continuous $l$\--adic representation $\tau: \Gal(\overline{K}/K)\to \GL_d(\mathcal{F})$ for which \begin{itemize} \item the image of $I$ is finite, \item any ($\Leftrightarrow$ every) choice of $\Phi_{\overline{K}/K}$ acts semisimply, \end{itemize} is of the form $$\tau \cong\bigoplus_{i}\rho_i\otimes \chi_i\quad\quad\quad\quad\mbox{ (for $\mathcal{F}$ sufficiently large),}$$ with \begin{itemize} \item $\rho_i$ Artin representations (with values in $\mathcal{F}$), \item $\chi_i$ $1$\--dimensional unramified: that is $\chi_i(I)=1$ and $\chi_i(\Phi_{\overline{K}/K})\in \mathcal{F}$. \end{itemize} \end{thm} It is not hard to show that for irreducible $l$\--adic representations with finite image of inertia, $\Phi_{\overline{K}/K}$ must act semisimply. \newpage \begin{exa} $\,$\\ \begin{minipage}{0.7\linewidth} Let us now consider an $l$\--adic representation for which the image of inertia $I=I_{\overline{K}/K}$ is infinite. Let $E$ be an elliptic curve over $K$, with $K/\Q_p$ finite, such that $E$ has split multiplicative reduction, and let $\rho_E:\Gal(\overline{K}/K)\to \GL_2(\Z_l)$ be the associated $l$-adic representation. Then $P(\rho_E, T)=1-T$, so there is a $1$\--dimensional $I$\--invariant space with trivial $\Phi_{\overline{K}/K}$ action: $$\rho_E(h)=\abcd{1}{?}{0}{?}\quad \forall h\in \Gal(\overline{K}/K).$$ Since $\det\rho_E=\chi_{cyc}^{-1}$, it follows that $\det\rho_E(h)=1$ for all $h\in I$, and so $\rho_E(h)=\abcd{1}{\ast}{0}{1}$ for $h\in I$ with $\ast \in \Z_l$ and not always zero (as otherwise there would exist a $2$\--dimensional inertia invariant subspace, which is not the case). \end{minipage} \begin{minipage}{0.3\linewidth} \xymatrix@dr@C=1pc{ \overline{\Q}_p \ar@/^2pc/@{-}[rrr]^{I}\ar@{-}[rr]& \, & L^{nr} \ar@{-}[r]^(0.3){\langle g\rangle}\ar@{-}[d]^{\Phi_L}& K^{nr}\ar@{-}[d] \\ \; & \;& L\ar@{-}[r]& K} \end{minipage} Hence, the image of inertia mod $l^n$ is of the form $$ \rho_E(I)\subseteq\left\{\abcd{1}{\ast}{0}{1}\in \GL_2(\Z/l^n\Z)\right\}\cong C_{l^n}. $$ As for every $n$ it is contained in a cyclic group of order $l^n$, the wild inertia (a $p$-group, $p\neq l$) maps to $\Id$, and the tame inertia $\Gal({K}^{\mbox{\scriptsize{tame}}}/K^{nr})\cong \prod_{p'\neq p}\Z_{p'}$ is such that only $\Z_l$ can act non\--trivially (and must do so). So, if $\Z_l=\langle g\rangle=\Gal(L/K^{nr})$, where $L=\cup_{n=1}^\infty K^{nr}({\sqrt[l^n]{\pi_K}})$ and $\Phi_L=\Phi_{L^{nr}/L}$, then $\rho_E$ factors through $\Gal(L^{nr}/K)$ and, with respect to a suitable $\Q_l$\--basis, $$ \rho_E(g)=\abcd{1}{1}{0}{1}\quad\quad \quad \mbox{and }\quad\quad\quad \rho_E(\Phi_L)=\abcd{1}{0}{0}{q}$$ where $q=\#\F_L$. More explicitly, $$\rho_E(h)=\abcd{1}{t_l(h)}{0}{\chi^{-1}_{cyc}(h)}\quad\mbox{for }h\in I,$$where $t_l:I\to\Z_l$ is the $l$\--adic tame character given by $h(\sqrt[l^n]{\pi_K})= \zeta_{l^n}^{t_l(h)}\sqrt[l^n]{\pi_K}$ for all $n$. \end{exa} \begin{rmk} Once we have $\rho_E(g)=\abcd{1}{t_l}{0}{1}$ and $\rho_E(\Phi_l)=\abcd{1}{?}{0}{?}$, the rest is, in fact, forced by the commutation relations of $\Phi_L$ and $I$. In particular, $\det\rho_E$ has to be $\chi_{cyc}^{-1}$. \end{rmk} \begin{dfn} The \emph{special representation} $sp(n)$ over $K$ is the $n$\--dimensional $l$-adic representation given by: $$sp(n)(h)= \left(\begin{matrix} 1 & t & t^2/2 & \cdots & t^{n{-}1}/(n{-}1)!\\ 0& 1 & t & \cdots & t^{n{-}2}/(n{-}2)!\\ & & \ddots& \ddots & \vdots\\ & & & & t\\ 0 & & & 0 & 1 \end{matrix}\right) \quad\mbox{ for }h\in I,$$ where $t=t_l(h)$ is the $l$\--adic tame character, and $$sp(n)(\Phi_{\overline{K}/K})= \left(\begin{smallmatrix} 1 & & & 0 \\ & {q} & & \\ & & \ddots& \\ 0 & & & q^{n{-}1}\end{smallmatrix}\right),\quad\mbox{ where }q=\#\F_K.$$ (Different choices of $\Phi_{\overline{K}/K}$ give rise to isomorphic representations.) Note that, in particular, $sp(1)$ is the trivial representation, and $sp(2)$ agrees with the one associated to an elliptic curve with split multiplicative reduction. \end{dfn} \begin{thm} Every continuous $l$\--adic representation $\tau: \Gal(\overline{K}/K)\to \GL_d(\mathcal{F})$ for which any ($\Leftrightarrow$ every) choice of $\Phi_{\overline{K}/K}$ acts semisimply on $\tau^{I'}$, for every $I'\subseteq I$ of finite index, is of the form $$\tau=\bigoplus_{i}\rho_i \otimes sp(n_i)\quad\quad\mbox{ (for $\mathcal{F}$ sufficiently large),}$$ where $n_i$ are integers and $\rho_i$ are continuous $l$\--adic representations with $\rho_i(I_{\overline{K}/K})$ finite and semisimple action of $\Phi_{\overline{K}/K}$. \end{thm} \begin{exa} Let $E/K$ be an elliptic curve with split multiplicative reduction. Then $\rho_E$ has infinite inertia image, so, by the theorem, $\rho_E=\rho\otimes sp(2)$ where $\rho$ is $1$\--dimensional with local polynomial $1-T$. Therefore, $\rho={\mathbf{1}}$ and $\rho_E(g)=sp(2)$. \end{exa} \bigskip \section{Independence of $l$} Throughout this section $K$ is a finite extension of $\Q_p$ and $l\neq p$ a prime. \begin{exa} The $l$\--adic cyclotomic character is defined such that $\chi_{cyc}(I)=1$ and $\chi_{cyc}(\Phi_{\overline{K}/K})=1/q$, where $q=\#\F_K$. Morally, this does not depend on $l$. However, $\Gal(\overline{K}/K)$ is a topological group, and limits do depend on $l$. For instance, the $\chi_{cyc}$ factors through $\cup_{n}K(\zeta_{l^n})/K$, which depends on $l$. \end{exa} \medskip \begin{exa} In the last lecture we had $E/\Q_5$ with $\rho_E(g)=\abcd{\zeta_3}{0}{0}{\zeta_3^{-1}}$ and $\rho_E(\Phi_L)=\abcd{0}{\sqrt{{-}5}}{\sqrt{{-}5}}{0}$. This representation does not depend on $l$ provided we only look at the inertia group and {\em integer} powers of $\Phi_L$. Non-integer powers of $\Phi_L$ again give problems because of different convergence properties in $\Q_l$ for different $l$. \end{exa} \medskip \begin{exa} Let $E/K$ be an elliptic curve with split multiplicative reduction. Then $\rho_E(g)=sp(1)$ which again, morally, does not depend on $l$. (However, the $l$\--adic tame character does). \end{exa} \medskip \begin{dfn} The \emph{Weil group} $W_{\overline{K}/K}$ is the subgroup of $\Gal(\overline{K}/K)$ consisting of elements whose image modulo $I_{\overline{K}/K}$ is an integer power of $\Phi_{\overline{K}/K}$, i.e.\ $$\xymatrix{ 0\ar[r]&I_{\overline{K}/K}\ar[r]\ar@{=}[d]& \Gal(\overline{K}/K)\ar[r]&\widehat{\Z}\cong\Gal(\F_{\overline{K}}/\F_K)\ar[r]&0\\ 0\ar[r]&I_{\overline{K}/K}\ar[r]&\;W_{\overline{K}/K}\;\ar[r] \ar@{}_{\rotatebox{90}{$\subseteq$}}[u]&\Z\cong\langle \Phi_{\overline{K}/K}\rangle\ar@{}^{\rotatebox{90}{$\subseteq$}}[u] \ar[r] &0. }$$ The topology of $W_{\overline{K}/K}$ is the same (profinite) one on $I_{\overline{K}/K}$ and discrete on $W_{\overline{K}/K}/I_{\overline{K}/K}$. \end{dfn} \begin{rmk} A continuous $l$\--adic representation over $K$ automatically restricts to a continuous representation $W_{\overline{K}/K}\to \GL_d(\mathcal{F})$. \end{rmk} \medskip \begin{thm} Let $E/K$ be an elliptic curve (or an abelian variety). Then the decomposition of $\rho_E$ as $\bigoplus_i \rho_i\otimes sp(n_i)$ is independent of $l$, i.e.\ $\rho_i\otimes_{\Q_l}\C$ are independent of $l$ as representations of $W_{\overline{K}/K}$, and the $n_i$ are independent of $l$. \end{thm} \begin{cor} Let $E/K$ be an elliptic curve (or an abelian variety) with potentially good reduction. Then $\rho_E\otimes_{\Q_l}\C$ is independent of $l$ as a representation of $W_{\overline{K}/K}$. \end{cor} \medskip \begin{cor} Let $E/K$ be an elliptic curve (or an abelian variety). The local polynomial $P(\rho_E,T)$ and the characteristic polynomials of $\rho_E(h)$ for every $h\in I_{\overline{K}/K}$ are independent of $l$ and lie in $\Q[x]$ (as they lie in $\Q_l[x]$ for all $l\neq p$). \end{cor} \medskip \begin{exa}Let $K/\Q_p$ be a finite extension with $p\geq 5$, and let $E/K$ be an elliptic curve with potentially good reduction. Then $\rho_E$ is tame (exercise $2$ of lecture $2$), so the action of $I_{\overline{K}/K}$ factors through a finite cyclic group $C_n=\langle g \rangle$ (as the tame inertia group is cyclic). By the above corollary, we must have $C_n=C_1, C_2,C_3, C_4$ or $C_6$ since the characteristic polynomial of $\rho_E(g)$ is quadratic and lies in $\Q[x]$. (Thus, in particular, $v_K(\Delta_E)\not\equiv 1,11\bmod 12$ and $E$ acquires good reduction over a quartic or a sextic totally ramified extension of $K$). \end{exa} \newpage \section*{Exercises to Lecture 3 \bigskip \setcounter{ex}{0} \begin{ex} Show that if $\rho_E$ is irreducible then $E$ acquires good supersingular reduction over a finite extension. Recall that $E/K$ has good supersingular reduction if $\#\tilde{E}(\F_K)\equiv 1 \bmod p$. \end{ex} \bigskip \bigskip \begin{ex}Show that an elliptic curve with potentially multiplicative reduction acquires split multiplicative reduction over a quadratic extension. \end{ex} \bigskip \bigskip \begin{ex}Decompose $sp(2)\otimes sp(2)$ as a direct sum of indecomposables. (If you like the representation theory of $\GL_2(\C)$ try also $sp(n)\otimes sp(m)$). \end{ex} \newpage
\section{Introduction} \label{sec:intro} \IEEEPARstart{M}{atching} pedestrians across multiple CCTV cameras have gained a lot of interest in recent years. Despite several attempts\cite{kviatkovsky2013color,custompict,bicovma2012} to address these challenges, it largely remains challenging mainly due to the following reasons. First, the images are captured under different lighting conditions. Therefore the perceived color of the subject appears to be different with respect to the illumination. Second, from surveillance cameras, no biometric aspects are available\cite{custompict}. Third, most often, the surveillance cameras will be of lower resolution\cite{kviatkovsky2013color}. Figure \ref{samplefigs} shows some examples of images from different datasets. Modern person re-identification systems primarily focus on two aspects. (1) A feature representation for the probe and gallery images and (2) a distance metric to rank the potential matches based on their relevance. In the first category, majority of the works has been done on designing low level features. Since each of the features capture different aspects of the images, usually a combination of these features are used to obtain a richer signature. In the second category, the person re-identification is formulated as a ranking or a metric learning problem. In this work, we focus on the feature representation aspect, specifically on color based features. Color based features have been proven to be an important cue for person re-identification\cite{kviatkovsky2013color}. An interesting insight on the importance of color features was demonstrated in an experiment conducted by Gray and Tao\cite{viewpointinvariant}. They used AdaBoost for giving weights to the most discriminative features. They observed that, over $75$ percent of the classifier weights were given to color based features. These observations support the fact that color has to be given much more attention than other handcrafted features based on shape, texture and regions. But due to the illumination variations across the camera views, the perceived color of same parts for a particular person appear to be different. Taking this observation into consideration and as validated from our experiments, we suggest that using color features {\it as it is}, i.e. RGB, HSV or YUV color histogram representation will not be adequate to achieve an illumination invariant representation. Hence we propose a data driven multilayer framework that learns invariant color features from raw pixel values as opposed to histogram or other color based handcrafted features. The proposed framework aims at learning an invariant representation for images from both camera views by transforming all the pixels to a color-constant space. In the color-constant space, the pixels are invariant to the camera and other environmental variables. Instead of modeling each of the variables, we exploit a data driven framework to explore the structures and patterns inherent in the image pixels. Feature learning approaches come in handy in this situation. We use an auto-encoder based framework to transform the 3-dimensional RGB pixel values to a higher dimensional space first and encode them using a dictionary. These encoded values are pooled over a region and concatenated to form the final representation of an image. This framework can be extended to a multilayer structure for learning complementary features at a higher level. Experiments were conducted on publicly available datasets such as ViPER\cite{viper}, Person Re-ID 2011\cite{hirzer11a} and CAVIAR4REID\cite{custompict}. From the results, it can be inferred that (1) by learning illumination invariant color features, significant improvement can be achieved in the results when compared with the traditional color histograms and other handcrafted features; and (2) when combined with other types of learned low-level and high-level features, it can achieve promising results in several benchmark datasets. The rest of this paper is organized as follows. Section 2 reviews some of the related works in color constancy, person re-identification and feature learning. Section 3 describes the motivation and the major contributions of this work. Section 4 describes the framework for learning the invariant color features. In section 5, we demonstrate the experimental evaluation of our method and compare with the other competing methods for person re-identification. In section 6, we perform an analysis of the obtained results and Section 7 concludes this paper. \section{Related Works} \subsection{Color Constancy} \label{subsec:cc} Human perceptual system has the ability to ensure that the perceived color of an object remains relatively constant even under varying illumination\cite{fastimplofccmorel2009}. Land and McCann proposed the Retinex theory\cite{retinexland71} to explain this perceptual effect. In one of the pioneering works\cite{forsythcolorconst}, Forsyth proposed the CRULE and MWEXT algorithms to achieve color constancy in Mondriaan world images by estimating the illuminant based on the information obtained from images such as reflectances and possible light sources. For a detailed overview of the color-constancy algorithms derived from the Retinex theory and \cite{forsythcolorconst}, we refer the reader to \cite{ccsurvey2011,ebner2007color,hordley2006scene}. All of the aforementioned works and the derived works are based on specific assumptions since the color constancy is an under-constrained problem \cite{forsythcolorconst, hordley2006scene}. For example, in \cite{forsythcolorconst}, the main assumption is constrained gamuts, i.e., the limited number of image colors which can be observed under a specific illuminant. Several other assumptions were based on the distribution of colors that are present in an image (e.g., white patch, gray-world and gray-edge). Majority of these works vary in their assumptions and therefore no color constancy algorithm can be considered as universal. Several works were done focusing on color features for object recognition. In one of the earliest works, Swain and Ballard\cite{swainballardcolorhistogram} identified that color histograms were stable representations over change in views. Gevers and Smeulders\cite{Geverscolorbasedobj} analysed different color spaces to achieve invariance to a substantial change in viewpoint, object geometry and illumination. But it was observed that the object recognition accuracy degrades substantially for all of the color spaces with a change in illumination color. A Diagonal Matrix Transform (DMT) is the basis of majority of the works\cite{berwicklee,healey1994global,Geverscolorbasedobj,kviatkovsky2013color} on color constancy. To improve the performance of DMT, spectral sharpening\cite{Finlayson94} derived for each camera can be incorporated. Berwick and Lee\cite{berwicklee} proposed a log-chromaticity color space to achieve specularity, illumination color and illumination pose invariance. A recent work\cite{kviatkovsky2013color} make use of the log-chromaticity color space to achieve invariant color features for person re-identification. The assumption in \cite{kviatkovsky2013color} is that the shape of the color cloud is sufficiently preserved in the log$\frac{R}{G}$, log$\frac{B}{G}$ space. But this assumption is violated in many real world images. The color cloud is formed based on sampled observations from upper body and lower part of the body. \figurename{\ref{fig_colorclouds}} shows some examples of those observations from the ViPER dataset. It can be seen that in a different view, the upper part of the body may have different colors for the same subject. In this paper, data driven techniques are used to discover a color constant space in contrast to methodologies based on strong assumptions. \begin{figure} \centering \includegraphics[width=0.1\linewidth]{11.jpg} \includegraphics[width=0.1\linewidth]{46.jpg} \includegraphics[width=0.075\linewidth]{173_0.jpg} \includegraphics[width=0.075\linewidth]{176_0.jpg} \includegraphics[width=0.1\linewidth]{person_0001a.png} \includegraphics[width=0.1\linewidth]{person_0002a.png} \\ \vspace{3pt} \includegraphics[width=0.1\linewidth]{12.jpg} \includegraphics[width=0.1\linewidth]{49.jpg} \includegraphics[width=0.075\linewidth]{173_90.jpg} \includegraphics[width=0.075\linewidth]{176_90.jpg} \includegraphics[width=0.1\linewidth]{person_0001.png} \includegraphics[width=0.1\linewidth]{person_0002.png} \caption{Some examples from CAVIAR4REID, ViPER and Person Re-ID 2011 datasets. Images clearly show the appearance changes due to the environmental variables - Illumination, Shading, Camera view angle. {\bf Best viewed in color}} \label{samplefigs} \end{figure} \subsection{Person Re-Identification} Person re-identification research has received a good amount of attention in recent years. As mentioned in section \ref{sec:intro}, existing works focus on the different steps that need to be taken for dealing with this problem. Majority of the works\cite{custompict,sdalf,bicovma2012,viewpointinvariant,zhao2013unsupervised} predominantly focus on the first step, i.e. designing features based on texture, color, shape, regions and interest points. Since the primary focus of this work is on color based feature design, a complete evaluation of all of the aforementioned features is beyond the scope of this paper. To obtain the global chromatic content, most of the works uses the color histogram features in RGB, HSV or YUV space. These color spaces do not possess the property of illumination invariance. In addition to a weighted HSV histogram, Maximally Stable Color Regions(MSCR) are also used in \cite{sdalf} to obtain the per-region color displacement. In a relatively closer work, Porikli\cite{btfporikli2003} and Javed\cite{btfjaved2005} proposed a Brightness Transfer Function(BTF) to find a transformation that maps the appearance of an object in one camera view to the other. But it should be noted that, the system has to be re-trained each time the illumination changes. In addition to that, the method adopts normalized histograms of object brightness values for BTF computation. Therefore, a pixel level correspondence cannot be achieved. It is important to note that all the aforementioned features for person re-identification are handcrafted focusing on specific cues. However, in this work we propose a model to learn color based features for a pair of camera to achieve the invariance properties across the two different views based on the intuition that the features should be invariant to change in environment and camera variables. To the best of our knowledge, this is the first work that focuses on a data driven learning of low-level color features. Since the framework is based on learning from the data, the scope of this work is beyond person re-identification. \subsection{Feature Learning} Recent researches have shown a growing interest in unsupervised feature learning methods such as auto-encoders\cite{vincent2008extracting}, sparse coding\cite{yang2009linear} and Deep Belief Nets\cite{hinton2006fast} since they are data driven and can be generalized to a larger extent. Since no handcrafted feature can be considered as universal, learning relations from data can be advantageous. Modeling complex distributions and functions have been a bottleneck in machine learning. Recent studies in deep learning indicate that such deep architectures can efficiently handle these challenges and have shown that better generalization can be obtained. Several successful algorithms have been proposed\cite{hinton2006fast},\cite{hinton2006reducing,lee2008sparse,bengio2007greedy,bengio2009learning} to train large networks such as deep belief networks and stacked auto-encoders. The intuition behind using large networks is that, to learn a complex function that computes the output from input, automatically learning features at multiple layers of abstraction can help to a large extend\cite{bengio2009learning}. Additionally, biological evidences substantiate that in the visual cortex, recognition happens at multiple layers\cite{serre2007quantitative}. However, all the above works focus on learning edges at different orientations in the first layer and higher level patterns in the further layers. In contrast to this, the proposed work addresses the problem of learning color features from data. We also propose a new method to learn a transformation and encoding simultaneously at a pixel level to achieve color constancy \begin{figure}[!t] \centering \subfloat[]{\includegraphics[trim = 28.57mm 112.83mm 114.73mm 106.36mm, clip,width=0.4\linewidth]{final.pdf} \label{fig_first_case}} \hfil \hspace{1cm} \subfloat[]{\includegraphics[trim = 114.3mm 112.83mm 27.432mm 106.36mm, clip,width=0.4\linewidth]{final.pdf} \label{fig_second_case}} \caption{Two examples to illustrate the difference in color clouds for the same person in different views due to appearance change in the second view. In (a), due to the presence of additional object(bag), the shape of the color clouds vary significantly. In (b), the color of the upper body has changed across the views which result in a different shape for the corresponding color clouds as shown in the plot. {\bf Best viewed in color}} \label{fig_colorclouds} \end{figure} \label{subsec:fl} \section{Motivation} \label{sec:motiv} Even though the importance of color features is proven in \cite{viewpointinvariant}, color features were not given much attention in person re-identification. Existing works use color {\it as it is}, i.e. computing the histogram in designed color spaces without any processing. However, varying lighting conditions affect the robustness of such designed features. These variations can be considered as noises that corrupt the actual pixel values. While designing color features, the performance will be affected unless these noises are handled well. The existing methods for achieving color constancy are based on strong assumptions about the statistics of color distribution, surfaces and its reflectance properties. Hence, a histogram representation in such a {\it weakly} corrected space will not be robust enough. Hand-engineered features focus on particular cues and adds more complexity to the system. Modeling each of the camera parameters and other environmental variables explicitly is practically impossible. Our intuition is that a {\it data driven} feature learning approach can discover a good intermediate representation from the input pixels. This means that, the relationship between the images across views can be learned by sampling observations from the images themselves. A robust representation should capture a certain amount of {\it information}, i.e. stable structures and patterns from the observed input. Though the inputs are corrupted, it can be reasonably assumed that there exists a space where the color patterns are invariant to these variables. A manifold learning interpretation of this problem is given in \cite{vincent2008extracting}. The corrupted inputs will lie away from the manifold and the objective is to learn a transformation to project them back to the manifold, in this case, the color constant space. With this assumption, we use a linear auto-encoder to discover such patterns from the data with the objective that in the transformed space, the representation for pixels of same colors should be as close as possible. Many researchers have empirically found that an encoding schemes for quantizing each local descriptors is essential for good performance. Sparse coding \cite{yang2009linear, yu2009nonlinear} has been found to achieve state-of-the-art performance for many classification problems. In this work, a sparse coding technique is adopted to encode each pixel by a set of dictionaries or codebooks. Previous works consider feature computation and encoding independently. However, these two approaches should be consistent with each other. Therefore, we propose a joint learning framework to obtain a transformation and the codebook simultaneously while enforcing the final encoded representation of each pixel belonging to the same color to be same. As validated from our experiments, we observe that the joint learning framework helps to obtain a robust representation and boost the performance significantly. Considering all the aforementioned challenges, we develop a data driven joint learning framework to handle these variables effectively. In summary, the contributions of our work are as follows \begin{figure*}[!t] \centering \subfloat[View A (Image and the Extracted Patch)]{\includegraphics[trim = 5mm 244mm 146mm 0mm, clip,width=0.2\linewidth]{sparsecodevshist.pdf} \label{fig_imga}} \hfil \subfloat[RGB Histogram]{\includegraphics[trim = 62.5mm 244mm 78mm 0mm, clip,width=0.2\linewidth]{sparsecodevshist.pdf} \label{fig_imgb}} \hfil \subfloat[Encoding using our approach]{\includegraphics[trim = 134mm 240mm 3mm 0mm, clip,width=0.2\linewidth]{sparsecodevshist.pdf} \label{fig_imgc}} \\ \vspace{-0.8cm} \subfloat[View B (Image and the Extracted Patch)]{\includegraphics[trim = 5mm 187mm 146mm 57mm, clip,width=0.2\linewidth]{sparsecodevshist.pdf} \label{fig_imgd}} \hfil \subfloat[RGB Histogram]{\includegraphics[trim = 62.5mm 187mm 78mm 57mm, clip,width=0.2\linewidth]{sparsecodevshist.pdf} \label{fig_imge}} \hfil \subfloat[Encoding using our approach]{\includegraphics[trim = 134mm 183mm 3mm 57mm, clip,width=0.2\linewidth]{sparsecodevshist.pdf} \label{fig_imgf}} \caption{(a) Training image from view A and a patch sampled from that image. (b) RGB histogram (216 dimensional) of the extracted patch. (c) Randomly sampled pixel encoded using our approach. (d) Training image from view B and a patch sampled from the corresponding region as patch from Image in (a). (e) RGB histogram (216 dimensional) of the extracted patch. (f) Encoding of a randomly sampled pixel from this patch. It can be seen from (b) and (e), the RGB histogram representation for the corresponding sampled patches are not close to each other. (c) and (f) shows the result of the encoding using our approach. It can be seen that, using our formulation, the obtained encoding is very close to each other for corresponding pixels in the sampled patch. We learn the transformation and Dictionary jointly so that the encoding is same for the two sampled patches. {\bf Best viewed in color}} \label{fig_sparsecodevshist} \end{figure*} \begin{itemize} \item We propose a novel data driven approach for learning inter camera illumination invariant color features from the pixels sampled from matching pair of images. In contrast with the previous works such as color histograms, normalized color spaces and BTF, our approach is more robust and represents the color features in an efficient and discriminative way. \item We propose a joint learning framework which solves the coupled problem of learning a linear auto-encoder transformation and a dictionary to encode the features. Previous works make use of an independent strategy to obtain features and encode them. \item We show that color as a single cue can bring a good performance that beats several hand-engineered features designed for person re-identification and when combined with other types of learned low-level and high-level features, it can achieve promising performance in several challenging datasets. \end{itemize} \section{Approach} As mentioned in the previous sections, the appearance of color changes across camera views due to a stark change in illumination. \figurename{\ref{fig_imga}} and \figurename{\ref{fig_imgd}} show an example of such changes in the appearance. It can be seen that the patches sampled, as shown in \figurename{\ref{fig_imga}} and \figurename{\ref{fig_imgd}}, appears to be of different color. Histogram of the sampled patches appears to be as shown in the \figurename{\ref{fig_imgb}} and \figurename{\ref{fig_imge}} respectively. Finlayson et al.\cite{Finlayson94} have shown that the diagonal model is an accurate model to achieve color-constancy for narrow-band(sensitive to single wavelength) imaging sensors. Practically, such sensors do not exist and the diagonal model is thus considered as an approximate model to correct the images for its illumination changes. Patches sampled from the images are pre-processed using L2-norm as mentioned in \cite{weakinvariancedrew1998}. We use a linear auto-encoder to transform the pixels into a rich higher dimensional space. These transformed pixel values are encoded using a sparse coding technique to attain more discriminative information. The objective behind the encoding is that the encoded values for corresponding pixels should be the same(or very close). To achieve consistency between the linear transformation and the encoding, we adopt a joint learning strategy to optimize the linear auto-encoder transformation and dictionary learning simultaneously. The parameters are updated alternatively to find the optimal mapping and encoding for the pixel values. More details are given in the subsequent sections. \subsection{Training Patch Collection} To train the system, patches were extracted from the training images manually. As a selection principle, we carefully choose patches so that sampled patches are distributed among different colors under varying illuminations. Background information were discarded and patches were cropped from corresponding parts of the same subject in two views where the color appears to be different. This is very important to create an optimal model and it can be seen from our experiments that this model based on these sampled patches gives a very good performance. From each of the training image pairs in each dataset, we sample 3 pairs of patches by following the criteria mentioned above. \figurename{\ref{fig_imga}} and \figurename{\ref{fig_imgd}} show an example of pair of patches sampled from the ViPER dataset. \subsection{Objective Formulation and Optimization} Intuitively, in a color constant space, the pixel values of same color should be same for both the images. Therefore, the objective of the proposed framework is to compute a representation so that the pixel values are close to the corresponding pixels in the matching image. This objective is formulated by making use of an auto-encoder and sparse encoding technique by enforcing the final encoded pixels of same color to have same values. As mentioned in section \ref{sec:motiv}, auto-encoder captures the stable structures and patterns in the data and projects it into a color constant space. At the same time, taking the performance into consideration, a sparse encoding is also applied to the descriptors to represent them compactly. Mathematically: \begin{equation} \label{eqn_test} \mathop\text{minimize }_{\bf W_1, W_2, D, \alpha_1, \alpha_2} {\ell_{ae} + \ell_{sc} + \varepsilon_{en} + \Omega}, \\ \end{equation} \begin{eqnarray} \ell_{ae} &=& \frac{1}{m} \left \| \left ( W_2'\left ( W_1' X + b_1 \right ) + b_2 \right ) - X\right \|^2_2 \\ \ell_{sc} &=& \frac{\beta}{m} \left \| \left ( W_1'X + b_1 \right ) - D\alpha \right \|^2_2 \\ \varepsilon_{en} &=& \frac{\gamma}{m} \left \| \alpha_1 - \alpha_2 \right \|^2_2 \\ \Omega &=& \lambda \left ( W_1^2 + W_2^2 \right )+ \rho\left \| D \right \|^2_2 + \frac{\eta}{m}\sqrt{\alpha^2 + \delta } \\\nonumber \end{eqnarray} where $\ell_{ae}$ is the auto-encoder loss function, $\ell_{sc}$ is the loss due to the sparse encoding, $\varepsilon_{en}$ is the error of the encoded values and $\Omega$ is the regularization term. $\Omega$ is essential to avoid learning trivial values and to enforce sparsity. Here, $X = [X^1 \quad X^2]\in\mathbb{R}^{3\times 2m}$, where $X^1 = [x^1_1,x^1_2,...,x^1_m]\in\mathbb{R}^{3\times m} $ and $X^2 = [x^2_1,x^2_2,...,x^2_m]\in\mathbb{R}^{3\times m}$ are the RGB values of $m$ randomly sampled pixels from patches extracted from view A images and the corresponding pixels from the patches extracted from view B images respectively. $W_1 \in \mathbb{R}^{3\times h}$ is the linear transformation matrix that transforms each of the pixels into a higher dimensional space and $b_1\in \mathbb{R}^{1\times h}$ is the bias term. Similarly, $W_2 \in \mathbb{R}^{h\times 3}$ is the transformation of the higher dimensional space into the original $3$ dimensional space and $b_2\in \mathbb{R}^{1\times 3}$ is the bias term. As mentioned in section \ref{subsec:cc}, previous works have proven that, a linear transformation is sufficient to transform images under an unknown illuminant to images under the canonical illuminant. We borrow this intuition into our work and therefore we use a linear auto-encoder. $D \in \mathbb{R}^{h\times d}$ are the basis vectors(Dictionary or Codebook) to encode each of the transformed pixel values in the higher dimensional space where $d$ is the number of such learned dictionaries. The encoding, $\alpha = [\alpha_1 \quad \alpha_2]\in\mathbb{R}^{d\times 2m} $ is sparse and is represented by $\alpha_1$ and $\alpha_2$ for $X^1$ and $X^2$ respectively. The optimization is done alternatively between $\alpha$, $D$ and $ (W_1,b_1,W_2,b_2)$. We use L-BFGS gradient based optimization procedure to update these values. The gradients with respect to each of the terms are given in Appendix A. Initially, $\alpha_1$ and $\alpha_2$ are updated based on their gradients while keeping $D$ and $ (W_1,b_1,W_2,b_2)$ fixed. Then $D$ is updated keeping $\alpha_1$, $\alpha_2$ and $(W_1,b_1,W_2,b_2)$ fixed. Finally, keeping $\alpha_1$, $\alpha_2$ and $D$ fixed, $(W_1,b_1,W_2,b_2)$ are updated together. For a simple and straightforward gradient based optimization, we replace $\left | \alpha \right |_1$, the L1 norm with an approximation that can smooth it at the origin, $\sqrt{\alpha^2 + \delta }$. $\delta$ is infinitesimally small ($1 \times 10^-4$). Theoretically, the gradient based optimization is simple for the above objective functions. But for faster convergence and a good optima, it requires a bit of finesse. With that in consideration, practically, good initializations are required for $W_1$, $W_2$ and $D$. We give the objective function for the initializations in Appendix B. Initialization helps to achieve faster convergence as well as a better local optima. Joint learning is done until convergence. During testing, for each pixel in an image, we first use $W_1$ and $b_1$ to transform it into a higher dimensional space and use the learned dictionary $D$ to compute the sparse codes. Therefore, the dimensionality of the features for each of the input image will be $ M \times N \times d$ where $ M \times N \times 3$ is the input image dimensions. \figurename{\ref{fig_imgc}} and \figurename{\ref{fig_imgf}} show the representation obtained by us using the joint learning framework. It can be seen that the error of encoding for the pixel sampled from the two patches is much less than the histogram representation of the patch. \subsection{Multi-layer Framework} The same formulation can be extended to a multilayer framework so that the representations are close to each other for the patches sampled from the images. To achieve this objective, we encode all the pixels in the sampled patches using the method mentioned in the above section and adopt a max pooling scheme over the $2\times 2$ regions to get the representation of a patch. The {\it max} pooling strategy makes the representation slightly translation invariant. We further adopt the same formulation with the linear transformation and dictionary learning and encode the observations over a patch and enforce the equality constraint over the corresponding samples from the image pairs. Once the representation is obtained in the second layer, max pooling is done over $4\times 4$ regions. For simplicity, we keep the dimensions $h$ and $d$ same for the second layer. We observed that further increasing the number of layers did not give any significant advantage over the performance. Also considering the complexity of the approach, we use only two layers for our model. \subsection{Parameters} The formulation contains several parameters such as the weights of the cost function terms, the dimensions of the linearly transformed space and the number of dictionary atoms. All the parameters were empirically determined by cross-validation. The obtained parameters are as follows. $h=60$, $d=250$, $\beta=1$, $\gamma=0.1$, $\lambda=3\times 10^-3$, $\rho=0.01$ and $\eta=0.01$. \section{Experiments} \label{sec:exp} We validate our algorithm on publicly available datasets such as ViPER\cite{viper}, Person Re-ID 2011\cite{hirzer11a} and CAVIAR4REID\cite{custompict}. The characteristics of these three datasets are ideal for the evaluation of the proposed Jointly Learned Color Features(JLCF) since the images were captured from two cameras under varying environments such as indoor and outdoor, bright and dark illumination and different view angles. Below, we list the baseline approaches we compare with. \begin{enumerate} \item {\bf Hist}: We compare our approach with the 3D histogram generated in RGB, HSV and YUV spaces. The histograms are computed from the images without any pre-processing for illumination changes. We use 6 bins for each of the channels so that the representation is 216 dimensional which is close to the 250 dimensional space proposed in this work. Image is divided into $8 \times 8$ blocks with a stride of 4 and for each of those blocks, we compute the histogram. We refer to the histogram representations as RGBHist, HSVHist and YUVHist. \item {\bf cHist}: cHist corresponds to the histogram of the images in the weakly corrected(L2 norm based correction) space. As mentioned for {\bf Hist}, the number of bins for each of the channels and the size of the image blocks are kept same. The representation based on the corrected color space is referred to as cRGBHist, cHSVHist and cYUVHist. \item {\bf $rg$Hist}: $rg$Hist corresponds to the histogram in the $rg$ space for the corrected images. $rg$ color channels are one of the first photometric invariant color channels proposed. The $rg$ space corresponds to \begin{equation} \label{eqn_rgspace} r = \frac{R}{R+G+B}, \qquad g = \frac{G}{R+G+B} \end{equation} The image is divided into blocks of $8\times 8$ and for each of the blocks, we compute histogram of 16 bins for each channel. The final representation for each block will be 256 dimensional. \item {\bf Opponent}: The opponent color space is invariant to specularity. It can be computed by \begin{equation} \label{eqn_opponent} \begin{aligned}[left] \begin{bmatrix} O^1\\ O^2\\ O^3 \end{bmatrix}=\begin{bmatrix} \frac{1}{\sqrt{2}} &\frac{-1}{\sqrt{2}} &0 \\ \frac{1}{\sqrt{6}} &\frac{1}{\sqrt{6}} &\frac{-2}{\sqrt{6}} \\ \frac{1}{\sqrt{3}} &\frac{1}{\sqrt{3}} &\frac{1}{\sqrt{3}} \end{bmatrix} \begin{bmatrix} R\\ G\\ B \end{bmatrix} \end{aligned} \end{equation} Image is divided into $8\times 8$ blocks and the final histogram representation is computed as mentioned in \cite{evaluatinggevers2010}. \item {\bf C}: The C color space adds photometric invariant with respect to shadow shading to the opponent color space. It is computed by normalizing Opponent descriptor by the intensity. \begin{equation} C = \begin{bmatrix} \frac{O^1}{O^3} &\frac{O^2}{O^3} &O^3 \end{bmatrix}^T \end{equation} For $8\times 8$ blocks, the histogram is computed as mentioned above for the opponent color space. \item {\bf Independent Learning}: The learning strategy will be optimizing the auto-encoder transformation first and then obtaining the sparse codes of the transformed pixel values without joint learning. The objective given in appendix A for initialization is used to find the optimal auto-encoder transformation. After obtaining the transformation, dictionary is learned using $\ell_{sc}$ in equation \ref{eqn_test}. \item {\bf JLCF without Color Constancy(JLCF WCC)}: To show that the learning based on color constancy is important, we develop the representation without the color constancy term, i.e., excluding the $\left \| \alpha_1 - \alpha_2 \right \|^2_2$ term in equation \ref{eqn_test}. \end{enumerate} The final representation of an image in each of the color space is obtained by concatenating the histogram of each blocks in the image. Keeping the settings same for all of the baselines, we use LADF\cite{ZhenliShiyu_CVPR2013} metric learning framework for all the comparisons. For baseline 6, 7 and the proposed JLCF, the matching score of first and second layer is combined to form the final score. The results are reported based on the Cumulative Matching Characteristics(CMC)\cite{cmcmoon2001computational}. Each of the datasets, experimental settings and their evaluations are given in detail in the following subsections. Since the color features are complementary to other types of learned low-level and high-level features, we also perform experiments by combining with them to compare with state-of-the-art results. The following features are combined with the proposed color features. \begin{enumerate} \item {\bf AE}: Single layer auto-encoder features are learned at a patch of size $8\times 8\times 3$. 400 filters are learned and the filter response of patches for an image are pooled over $8\times 8$ regions. Vectorizing this representation gives the final feature for a single image. The features learned from a single layer auto-encoder are gabor-like edges. \item {\bf CNN}: The imagenet pre-trained model of Caffe\cite{Jia13caffe} which follows the architecture in \cite{krizhevsky2012imagenet} is used to obtain high-level features. The dimensionality of the obtained feature is $4096$. \end{enumerate} \begin{table}[!t] \renewcommand{\arraystretch}{1.3} \caption{Performance Comparison of different baselines and photometric invariant color spaces on the Standard Evaluation Split of the ViPER, Person Re-ID 2011 and CAVIAR4REID Datasets. Proposed Jointly Learned Color Features(JLCF) outperform all the baselines.} \label{table_Base} \centering \subfloat[ViPER]{ \begin{tabular}{|c|c|c|c|c|c|} \hline \bfseries Method & \bfseries Rank 1 & \bfseries Rank 5 & \bfseries Rank 10 & \bfseries Rank 15 \\\hline\hline RGBHist & 7.59 \% & 27.53 \% & 43.35 \% & 54.74 \% \\ \hline HSVHist & 11.07 \% & 33.23 \% & 48.42 \% & 60.44 \%\\ \hline YUVHist & 1.90 \% & 10.44 \% & 18.35 \% & 24.05 \%\\ \hline cRGBHist & 7.59 \% & 27.53 \% & 43.67 \% & 54.74 \%\\ \hline cHSVHist & 12.34 \% & 38.29 \% & 52.53 \% & 65.19 \%\\ \hline cYUVHist & 3.48 \% & 13.60 \% & 22.46 \% & 28.16 \%\\ \hline rgHist & 1.90 \% & 7.59 \% & 16.46 \% & 22.78 \% \\ \hline C Color Space & 7.59 \% & 26.27 \% & 40.50 \% & 48.73 \%\\ \hline Opponent & 7.91 \% & 25.31 \% & 38.29 \% & 43.67 \%\\ \hline {\bf \begin{tabular}{@{}c@{}}Independent \\ Learning\end{tabular}} & 20.25 \% & 47.78 \% & 64.24 \% & 76.89 \%\\ \hline {\bf JLCF WCC} & 19.94 \% & 48.73 \% & 65.50 \% & 75.00 \%\\ \hline {\bf Proposed JLCF} & {\bf 26.27 \%} & {\bf 51.90 \%} & {\bf 67.09 \%} & {\bf 78.17 \%}\\ \hline \end{tabular} } \subfloat[Person Re-ID 2011]{ \begin{tabular}{|c|c|c|c|c|c|} \hline \bfseries Method & \bfseries Rank 1 & \bfseries Rank 10 & \bfseries Rank 20 & \bfseries Rank 50 \\\hline\hline RGBHist & 1 \% & 14 \% & 25 \% & 41 \%\\ \hline HSVHist & 0 \% & 7 \% & 14 \% & 17 \%\\ \hline YUVHist & 1 \% & 6 \% & 12 \% & 27 \%\\ \hline cRGBHist & 1 \% & 13 \% & 25 \% & 42 \%\\ \hline cHSVHist & 0 \% & 11 \% & 22 \% & 31 \%\\ \hline cYUVHist & 3 \% & 9 \% & 14 \% & 33 \%\\ \hline rgHist & 0 \% & 2 \% & 6 \% & 21 \% \\ \hline C Color Space & 3 \% & 12 \% & 20 \% & 33 \% \\ \hline Opponent & 3 \% & 21 \% & 30 \% & 42 \% \\ \hline {\bf \begin{tabular}{@{}c@{}}Independent \\ Learning\end{tabular}} & 6 \% & 32 \% & 41 \% & 62 \%\\ \hline {\bf JLCF WCC} & 9 \% & 36 \% & 47 \% & 60 \%\\ \hline {\bf Proposed JLCF} & {\bf 14 \%} & {\bf 45 \%} & {\bf 55 \%} & {\bf 72 \%}\\ \hline \end{tabular} } \subfloat[CAVIAR4REID]{ \begin{tabular}{|c|c|c|c|c|} \hline \bfseries Method & \bfseries Rank 1 & \bfseries Rank 5 & \bfseries Rank 10 & \bfseries Rank 20 \\\hline\hline RGBHist & 19.79 \% & 65.47 \% & 81.89 \% & 98.94 \% \\ \hline HSVHist & 18.73 \% & 56.42 \% & 76.00 \% & 93.68 \% \\ \hline YUVHist & 17.05 \% & 51.36 \% & 68.84 \% & 94.52 \% \\ \hline cRGBHist & 24.21 \% & 68.42 \% & 84.84 \% & 97.47 \% \\ \hline cHSVHist & 20.21 \% & 59.36 \% & 77.47 \% & 94.94 \%\\ \hline cYUVHist & 22.73 \% & 53.26 \% & 72.42 \% & 96.84 \% \\ \hline rgHist & 12.84 \% & 43.36 \% & 68.84 \% & 92.63 \%\\ \hline C Color Space & 22.52 \% & 56.63 \% & 72.21 \% & 94.73 \%\\ \hline Opponent & 25.05 \% & 58.10 \% & 71.78 \% & 90.52 \%\\ \hline {\bf \begin{tabular}{@{}c@{}}Independent \\ Learning\end{tabular}} & 25.47 \% & 66.52 \% & 86.526 \% & 99.36 \%\\ \hline {\bf JLCF WCC} & 26.68 \% & 65.47 \% & 82.73 \% & 98.31 \%\\ \hline {\bf Proposed JLCF} & {\bf 32.63 \%} & {\bf 67.15 \%} & {\bf 87.57 \%} & {\bf 99.36 \%} \\ \hline \end{tabular} } \end{table} \subsection{ViPER Dataset} ViPER\cite{viper} is the most popular and challenging dataset to evaluate Person Re-Identification. The dataset contains 632 pedestrians from arbitrary viewpoints under varying illumination conditions and have relatively low resolution. The images are normalized to a size of $128\times 48$. We use the same settings as mentioned in \cite{zhao2013unsupervised,ZhenliShiyu_CVPR2013} for the evaluation. Table \ref{table_Base} shows the performance comparison of our approach with different baseline methods and photometric invariant color spaces. Experimental results suggest that the encoding based on the joint learning helps to a very large extend in achieving good performance. Table \ref{table_Base} also shows our comparison with the independent learning strategy as well as the joint learning framework without the color constancy term, $\varepsilon_{en}$ in equation \ref{eqn_test}. An evaluation of the state of the art algorithms is given in table \ref{table_viperState}. The \cite{ZhenliShiyu_CVPR2013}, \cite{zhang2014novel} and \cite{zhang2014structured} out performs our methods at different ranks. We observed that this is due to the lack of an optimal score combining mechanism for color and texture features. \cite{Zhao_2014_CVPR} achieved state-of-the-art results by combining with LADF. However, using color as a single cue, we achieve comparable results with several state-of-the-art methods based on multiple handcrafted features. \subsection{Person Re-ID 2011} The Person Re-ID 2011 dataset\cite{hirzer11a} consists of images extracted from multiple person trajectories recorded from two different, static surveillance cameras. Images from these cameras contain a sheer difference in illumination, background, camera characteristics and a significant view point change. Multiple images per person are available in each camera view. There are 385 person trajectories from one view, 749 from the other and 200 people appear in both views. For more details regarding the dataset, we refer the reader to \cite{hirzer11a} and \cite{hirzer2012relaxed}. In our experiments we do a multi-shot re-id with the same settings as mentioned in \cite{hirzer2012relaxed} and compare our results with the baselines as well as the state-of-the-art results. Table \ref{table_Base} shows the comparison of our approach with the baselines and photometric invariant color spaces. Table \ref{table_viperState} shows the comparison of our approach with state-of-the-art results in Person Re-ID 2011 dataset. As shown in the results, our method clearly outperforms the baselines, different photometric invariant color spaces and when combined with AE and CNN features, it outperforms the state-of-the-art results. \subsection{CAVIAR4REID} CAVIAR4REID\cite{custompict} is a person re-id evaluation dataset which was extracted from the well known caviar dataset for evaluation of people tracking and detection algorithms. It is a relatively smaller dataset which contains a total of 72 pedestrians(50 of them in both camera views and the remaining 22 with one camera only) taken from a shopping mall in Lisbon. Re-identification in this dataset is challenging due to a large variation in the resolution, illumination, occlusion and pose changes. The experimental settings are kept the same as in \cite{kviatkovsky2013color}. Table \ref{table_Base} shows that the illumination invariant signatures are performing significantly better than the baseline color features and different photometric invariant color spaces. We also compare with other standard approaches by combining JLCF with AE and CNN features and the results are reported in table \ref{table_viperState}. Similar to the ViPER dataset, we observed that lack of an optimal score combining mechanism affects the performance at rank 1. However, it should be noted that we achieve the best results at higher ranks. \begin{table}[!t] \renewcommand{\arraystretch}{1.3} \caption{Performance Comparison of different state-of-the-art results on the Standard Evaluation Split of the ViPER, Person Re-ID 2011 and CAVIAR4RED Datasets. Proposed Jointly Learned Color Features(JLCF) combined with AE and CNN can achieve promising results in the following datasets.} \label{table_viperState} \centering \subfloat[ViPER]{ \begin{tabular}{|c|c|c|c|c|} \hline \bfseries Method & \bfseries Rank 1 & \bfseries Rank 5 & \bfseries Rank 10 & \bfseries Rank 15 \\\hline\hline SDALF\cite{sdalf} & 19.87 \% & 38.89 \% & 49.37 \% & 58.46 \% \\ \hline CPS\cite{custompict} & 21.84 \% & 44.00 \% & 57.21 \% & 65.18 \% \\ \hline eBiCov\cite{bicovma2012} & 20.66 \% & 42.00 \% & 56.18 \% & 63.11 \% \\ \hline ELF\cite{viewpointinvariant} & 12 \% & 41.50 \% & 59.50 \% & 68.00 \% \\ \hline SalMatch\cite{zhao2013person} & 30.15 \% & 52.31 \% & 65.53 \% & 73.41 \%\\ \hline PatMatch\cite{zhao2013unsupervised} & 26.90 \% & 47.46 \% & 62.34 \% & 73.41 \%\\ \hline \begin{tabular}{@{}c@{}} Mid-level \\Features\cite{Zhao_2014_CVPR}\end{tabular} & 29.11 \% & 52.34 \% & 65.95 \% & 73.92 \%\\ \hline LADF\cite{ZhenliShiyu_CVPR2013} & 29.43 \% & 63.29 \% & 76.27 \% & 83.23 \% \\ \hline VWCM\cite{zhang2014novel} & 30.70 \% & 62.97 \% & 75.95 \% & 81.01 \% \\ \hline PRSP\cite{zhang2014structured} & 38.92 \% & 67.41 \% & 80.38 \% & 84.81 \% \\ \hline {\bf Proposed JLCF} & 26.27 \% & 51.90 \% & 67.09 \% & 78.17 \%\\ \hline {\bf\begin{tabular}{@{}c@{}} Proposed JLCF + \\ AE + CNN\end{tabular}} & 32.28 \% & 56.02 \% & 70.26 \% & 79.75 \% \\ \hline \end{tabular} } \\ \subfloat[Person Re-ID 2011]{ \begin{tabular}{|c|c|c|c|c|} \hline \bfseries Method & \bfseries Rank 1 & \bfseries Rank 10 & \bfseries Rank 20 & \bfseries Rank 50 \\\hline\hline Descr. Model\cite{hirzer11a} & 4 \% & 24 \% & 37 \% & 56 \% \\ \hline RPML\cite{hirzer2012relaxed} & 15 \% & 42 \% & 54 \% & 70 \%\\ \hline {\bf Proposed JLCF} & 14 \% & 45 \% & 55 \% & 72 \%\\ \hline {\bf\begin{tabular}{@{}c@{}} Proposed JLCF + \\ AE + CNN\end{tabular}} & {\bf 19 \%} & {\bf 47 \%} & {\bf 57 \%} & {\bf 75 \%}\\ \hline \end{tabular} } \\ \subfloat[CAVIAR4REID]{ \begin{tabular}{|c|c|c|c|c|c|} \hline \bfseries Method & \bfseries Rank 1 & \bfseries Rank 5 & \bfseries Rank 10 & \bfseries Rank 20 \\\hline\hline HPE\cite{Bazzani2012898} & 9.7 \% & 33.2 \% & 55.6 \% & 76.3 \%\\ \hline LF\cite{pedagadilfda} & 36.1 \% & 51.2 \% & 88.6 \% & 97.5 \%\\ \hline LADF\cite{ZhenliShiyu_CVPR2013} & 29.64 \% & 62.01 \% & 78.52 \% & 94.23 \%\\ \hline SSCDL\cite{Liu_2014_CVPR} & {\bf 49.1}\% & 80.2 \% & 93.5 \% & 97.9 \%\\ \hline {\bf Proposed JLCF} & 32.63 \% & 67.15 \% & 87.57 \% & 99.36 \% \\ \hline {\bf\begin{tabular}{@{}c@{}} Proposed JLCF + \\ AE + CNN\end{tabular}} & 45.89 \% & {\bf 80.84 \%} & {\bf 94.10 \%} & {\bf 100.00 \%}\\ \hline \end{tabular} } \end{table} \section{Analysis} In this section, we give the analysis of our approach and compare it with the baseline methods for color features. \subsection{JLCF vs Designed color spaces} Color histograms are representations which capture the color distribution. However, the difference in illumination can cause a significant change in the appearance which is caused by the variation in the RGB pixel values. Therefore, without the illumination correction, the histogram will not be a robust representation for color images. Using L2-norm, we do a correction for each of the images and then obtain the histogram in such a corrected space, the {\bf cHist}. Our approach uses illumination correction and find an optimal transformation to encode the pixel values in such a way that pixels corresponding to similar colors are close enough. The other photometric invariant color spaces can be considered as handcrafted features addressing specific cues as mentioned in section \ref{sec:exp}. Table \ref{table_Base} shows the comparisons of our method with the baseline approaches and it can be seen that JLCF clearly outperforms all of them. This is due to the fact that the L2 norm based correction which is inspired from the diagonal model is a weak illumination correction due to the strong assumptions. The comparison also shows that data driven learning approach is better than handcrafted color features. \subsection{JLCF vs Independent Learning} As mentioned in section \ref{sec:motiv}, to be consistent with each other, the linear auto-encoder transformation and dictionary for sparse coding must be learned jointly. As it can be seen from the results in table \ref{table_Base}, the joint learning improves the performance significantly for all the datasets. This is due to the fact that, for the encoding of each pixel, an optimal dictionary which can give same representation for pixels of same color has to be learned together with the linear transformation. \subsection{Proposed JLCF vs JLCF without Color Constancy} The objective of our approach is that for similar colors, the encoding must be same. Since the pixel values are corrupted by varying lighting conditions, this cannot be achieved by merely taking a histogram in pre-defined color spaces or sparse coding unless the pixel values are {\it close enough}. With this objective, the formulation in equation \ref{eqn_test} encodes the color information in such a way that for same colors, the encoding error is minimized. We conduct experiments with and without the color constancy term ($\varepsilon_{en}$) and report the results in table \ref{table_Base}. It can be seen that the color constancy based encoding improves the performance significantly. This is due to the fact that, without the color constancy term, the objective function merely encodes the pixels in a new space without any correction for the varying lighting conditions. But it should also be noted that the sparse encoding of linearly transformed pixel values results in a much better representation than histograms. \section{Conclusion} In this paper, we propose a novel data driven framework for learning color features to handle illumination and other lighting condition changes across two camera views. In contrast to the previous works based on auto-encoders and sparse coding, we combine them to learn a robust encoding jointly by forcing similar colors to be close to each other. We have also evaluated several baseline methods for achieving color constancy and have shown superior performance over all of them. By combining with other types of learned low-level and high-level features, we achieve promising results in several benchmark datasets. \ifCLASSOPTIONcaptionsoff \newpage \fi \bibliographystyle{IEEEtran}
\section{Introduction} We propose a new computer assisted method for rigorously proving invertibility and bounding the norm of the inverse of operators of the form\footnote{Conditions on $A(t)$ and the precise functions spaces in which we work will be specified later.} \begin{equation}\label{lin_bvp} F[v](t)=\left(v(t)-v(0)-\int_0^t A(s)v(s)ds,B_0v(0)+B_1v(1)\right),\hspace{2mm} A(t),B_i\in\mathbb{R}^{n\times n}. \end{equation} Operators of this form correspond to linear two point boundary value problems (BVPs) and arise naturally as the Fr\'{e}chet derivative of operators defining nonlinear BVPs. Bounds on the norm of the inverse of operators of this type are required in a posteriori proofs of existence and local uniqueness of nonlinear equations via the Newton-Kantorovich theorem~\cite{kantorovich1964functional}. Computer assisted methods have been successfully applied to many problems in differential equations and dynamical systems, including chaos in the R\"{o}ssler \cite{0951-7715-10-1-016} and Lorenz \cite{Galias1998165} systems, the existence of the Lorenz attractor \cite{Tucker19991197}, flow (in)stability \cite{Asen11981,watanabe2009computer}, existence of heteroclinic~\cite{Lessard2014}, homoclinic~\cite{doi:10.1137/100812008,doi:10.1137/12087654X,Arioli201551}, and periodic~\cite{Coomes,radii_poly} orbits, solution of the Feigenbaum equation~\cite{koch}, and the (in)stability of matter~\cite{fefferman1986relativistic}. Generalizations to elliptic and parabolic PDEs are discussed in, for example, \cite{Nakao1991323,nagatou1999approach,nakao2001numerical,plum2008existence}. As it pertains to two point boundary value problems, computer assisted proofs are typically based on a fixed point theorem \cite{Yamamoto,Plum,Nakao1992489}. In particular, various forms of the Newton-Kantorovich theorem were used in \cite{McCarthy,Kedem,Takayasu}. A method based on differential inequalities as presented in \cite{Gohlen}, a method using radii polynomials is given in~\cite{radii_poly}, and \cite{doi:10.1137/13090883X} utilized Chebyshev series. The methods in \cite{McCarthy,Gohlen,Plum,Takayasu,Nakao1992489,Fogelklou20111227,Fogelklou2012} focused on second order equations; \cite{McCarthy} worked with an equivalent integral formulation while \cite{Takayasu,Nakao1992489} used a finite element based method. \cite{Fogelklou2012} treated an integral boundary condition. The results in \cite{Urabe,Kedem} utilized the theory of initial value problem (IVP) fundamental solutions to treat $n$-dimensional two point BVPs, the former focusing on the periodic case. Our method most closely resembles that of Ref.~\cite{Kedem} and applies to $n$-dimensional systems without any special assumptions on the form of the vector field. The present method differs from \cite{Kedem} in the use of the BVP fundamental solution and Green's function as the primary tool, as opposed to the IVP fundamental solution. See also \cite{doi:10.1137/140987973}, which used a Green's function method to solve a singular problem on an unbounded domain. A general outline of our method is as follows. \begin{enumerate} \item Generate (non-rigorous) approximations, $\tilde \Phi_j$, $\tilde G_{i,j}$ to the BVP fundamental solution and Green's function, respectively, on a mesh. \item Extend these to piecewise polynomial approximations $\tilde\Phi(t)$ and $\tilde G(t,s)$ on the whole domain. \item Use $\tilde \Phi$ and $\tilde G$ to define an approximate inverse, $H$, to the BVP operator $F$. \item Using estimates involving $H$, prove existence of an exact inverse to $F$ and derive a machine computable formula that bounds the norm of $F^{-1}$. \end{enumerate} The use of the BVP fundamental solution makes the method presented here applicable in a range of situations for which IVP based methods are ill suited, namely when the IVP fundamental solution or its inverse has a `large' norm but the corresponding BVP fundamental solution and Green's function has a `moderately sized' norm. This is often the case in singularly perturbed problems and so the effectiveness of the method is demonstrated on two such examples in section \ref{sec:num_tests}. We will refer to these imprecisely defined `large' and `moderate' regimes as unstable and stable respectively, similar to the discussion in \cite{ascher_petzold}. To obtain completely rigorous results, some means of bounding the floating point rounding error is needed, typically based on the theory of interval arithmetic. There are many references from which to learn more about validated numerics and interval arithmetic, for example~\cite{moore1979methods,kearfott1996interval,jaulin2001applied,moore2009introduction,tucker2011validated}. Results in this paper are presented so that if interval arithmetic is used for all arithmetic operations and interval valued extensions are available for the vector field of the differential equation and its derivatives then the resulting computer generated existence proofs and error bounds are mathematically rigorous (we use the MATLAB package INTLAB \cite{Ru99a} to obtain rigorous bounds). However, in light of the additional computational cost of interval arithmetic as compared to floating point, if complete rigor is not needed then the method can just as easily be used with traditional floating point operations to obtain approximate error bounds. In section \ref{sec:newton} we fix some notation and discuss a general strategy for a posteriori existence proofs via the Newton-Kantorovich theorem. For comparison, we outline the method of \cite{Kedem} in section \ref{sec:IVP_method}. Development of our method begins in section \ref{sec:gen} and continues in subsequent sections. Background on the Green's function of a BVP is given in section \ref{sec:greens}. The method is tested on several sample problems in section \ref{sec:num_tests}. \section{A Posteriori Existence Proofs via Newton-Kantorovich}\label{sec:newton} Consider a general non-linear two point BVP on $[0,1]$, \begin{equation}\label{nonlin_bvp} y^\prime(t)=f(t,y(t)),\hspace{2mm} g(y(0),y(1))=0, \end{equation} where $f:\mathbb{R}^{n+1}\rightarrow\mathbb{R}$ is continuous, differentiable in $y$, $D_yf$ is continuous on $\mathbb{R}^{n+1}$, and $g:\mathbb{R}^{2n}\rightarrow\mathbb{R}^n$ is $C^1$. In particular, local existence and uniqueness of the IVP is guaranteed. In order to apply functional analytic methods to this problem it is convenient to follow \cite{Kedem} and define the Banach spaces \begin{equation}\label{X_def} X_1=C([0,1],\mathbb{R}^n), \hspace{2mm} C_0([0,1],\mathbb{R}^n)\equiv \{y\in X_1:y(0)=0\},\hspace{2mm} X_2=C_0([0,1],\mathbb{R}^n)\times\mathbb{R}^n \end{equation} where the $\infty$-norm is used on $\mathbb{R}^n$, denoted by $|\cdot|$, the sup-norm $\|y\|\equiv \sup_{[0,1]}|y(t)|$ is used on the function spaces, and the norm $\|(y,b)\|=\max\{\|y\|,|b|\}$ is used on $X_2$. In this setting, the BVP (\ref{nonlin_bvp}) can be put in a form more amenable to analysis as follows. Define the map $G:X_1\rightarrow X_2$ \begin{equation}\label{G_def} G[y](t)=\left(y(t)-y(0)-\int_0^tf(s,y(s))ds,g(y(0),y(1))\right). \end{equation} Solutions of the BVP are in one to one correspondence with zeros of $G$. Its first component, valued in $C_0([0,1],\mathbb{R}^n)$, is denoted by $G^1$ and its second component, valued in $\mathbb{R}^n$, by $G^2$. As shown in~\cite{Kedem}, $G$ is Fr\'{e}chet-$C^1$ with derivative at $y_0\in X_1$ given by \begin{align}\label{DF_formula} DG^1(y_0)[v](t)=&v(t)-v(0)-\int_0^tD_y f(s,y_0(s))v(s)ds,\\ DG^2(y_0)[v](t)=&D_{y_1}g(y_0(0),y_0(1))v(0)+D_{y_2}g(y_0(0),y_0(1))v(1).\notag \end{align} Having reformulated the problem in terms of finding zeros of a $C^1$ map between Banach spaces, a natural tool is the Newton-Kantorovich theorem; see Appendix \ref{newton_th} for a precise statement. Invertibility of $DG(y_0)$ along with a bound on the norm of its inverse are required to apply the theorem and these are typically the most difficult ingredients to obtain in practice. Once this has been done the remainder of the estimates are relatively straightforward and have been discussed in, for example, \cite{Kedem}. For this reason we will predominately focus on the invertiblility of $DG(y_0)$ i.e. on a posteriori methods for proving the existence and uniqueness to solutions of linear BVPs. \section{IVP Fundamental Solution Method}\label{sec:IVP_method} Motivated by the previous section, consider a linear BVP on $[0,1]$ as given by the bounded linear operator $F:X_1\rightarrow X_2$ defined by the formula (\ref{lin_bvp}), where $A(t)$ is a continuous $\mathbb{R}^{n\times n}$ valued map. In Theorem 1 of \cite{Kedem}, the author shows how the solvability of $F[v]=(r,w)$ is related to the matrix of fundamental solutions, defined as the solution to the initial value problem (IVP) \begin{equation*} Y^\prime(t)=A(t)Y(t),\hspace{2mm}Y(0)=I. \end{equation*} In our notation, this result is the following. \begin{theorem}\label{thm:ivp} Let $A(s)$ and $r(t)$ be continuous matrix and vector valued respectively with $r(0)=0$ and consider the two point BVP \begin{equation*} v(t)=v(0)+\int_0^t A(s)v(s)ds+r(t),\hspace{2mm} B_0v(0)+B_1v(1)=w. \end{equation*} Let $Y(t)$ be the corresponding matrix of fundamental solutions and define $R=B_0+B_1Y(1)$. Then the BVP has a unique solution iff $R$ is nonsingular and the solution is given by \begin{align*} v(t)=&Y(t)\bigg[R^{-1}B_0\int_0^tY^{-1}(s)A(s)r(s)ds\\ &+(R^{-1}B_0-I)\int_t^1Y^{-1}(s)A(s)r(s)ds-R^{-1}(B_1r(1)-w)\bigg]+r(t).\notag \end{align*} \end{theorem} It was also shown in \cite{Kedem} how to apply this theorem to rigorously prove solvability of the BVP and derive guaranteed error bounds using numerically constructed approximations to $Y(t)$ and $Y^{-1}(t)$. Because of the use of $Y$, this will be called the IVP fundamental solution method or simply the IVP method. A limitation of this method is that the stability properties of an IVP take center stage; the quantities $|Y(t)|$ and $|Y^{-1}(t)|$ feature critically in the error estimates derived in \cite{Kedem}. Unfortunately there are situations in which a BVP of interest is stable whereas the corresponding IVP is not. In such situations the bounds obtained by the IVP fundamental solution method will be very pessimistic due to the large size of $|Y(t)|$ and/or $|Y^{-1}(t)|$. As discussed in \cite{ascher_petzold}, perhaps the simplest test problem that exhibits these features is the BVP \begin{equation}\label{test_prob} y^{\prime\prime}(t)=y(t)\hspace{2mm} y(0)=1,\hspace{2mm} y(b)=0. \end{equation} The corresponding IVP fundamental solution matrix and its inverse are \begin{equation*} Y(t)=\left( \begin{array}{cc} \cosh(t) & \sinh(t) \\ \sinh(t) & \cosh(t)\\ \end{array} \right),\hspace{2mm} Y^{-1}(t)=\left( \begin{array}{cc} \cosh(t) & -\sinh(t) \\ -\sinh(t) & \cosh(t)\\ \end{array} \right). \end{equation*} These have exponentially increasing norms as a function of $b$. This is in spite of the fact that the value and derivative of the exact solution \begin{equation*} y(t)=\cosh(t)-\cosh(b)\sinh(t)/\sinh(b) \end{equation*} have $\mathcal{O}(1)$ norm. Even if one tries to avoid separately bounding the norms of $Y$ and $Y^{-1}$ and take advantage of the cancellation that can occur in $Y(t)R^{-1}B_0Y^{-1}(s)$ and $Y(t)(R^{-1}B_0-I)Y^{-1}(s)$, the above example is sufficient to show that this cancellation does not in general occur over the entire $(t,s)$ domain; the $1,1$ component of the former quantity is given by \begin{equation*} (Y(t)R^{-1}B_0Y^{-1}(s))^1_1=(\cosh(t)-\cosh(b)\sinh(t)/\sinh(b))\cosh(s) \end{equation*} which diverges at $t=0$, $s=b$ as $b\rightarrow\infty$. Even if cancellation does occur, it is difficult to compute $Y(t)$ and $Y^{-1}(s)$ to high enough accuracy that one could take advantage of such cancellation. In summary, while the IVP fundamental solution method suffices for problems whose IVP is stable, if one is interested in BVPs whose corresponding IVP is unstable then an alternative method is needed. Our contribution is designed to address such situations by utilizing the BVP fundamental solution and Green's function in place of the IVP fundamental solution. \section{A More General Framework}\label{sec:gen} We are now in position to begin discussing an alternative method for producing a posteriori existence proofs and error bounds based on the BVP fundamental solution. This method will be called the BVP fundamental solution method, or BVP method for short, in contrast to the IVP method. The BVP fundamental solution is introduced in the following section, but first we discuss the function spaces that will be used in our analysis. The input to any a posteriori method consists of a certain (numerically generated) approximate solution. For us this will consist of function values on a mesh $0=t_0<...<t_N=1$. Before error bounds can be produced, these need to be extended to an approximate solution on the entire interval $[0,1]$. While one could work in spaces of continuous functions by interpolating the values at the nodes, as done in the IVP method, we will find things more convenient (and the resulting error bounds more transparent) when this assumption is removed, both from the solution space and the matrix $A(t)$. In this section we outline the slightly generalized framework in which the equations will be formulated. To this end, fix a mesh, $t_i$, as above and let $Y_1$ be the space of functions that are piecewise continuous on this mesh. More precisely, $Y_1$ is the direct sum \begin{equation}\label{Y1_def} Y_1=\bigoplus_{j=1}^NC([t_{j-1},t_j],\mathbb{R}^n). \end{equation} On each factor the norm will be taken to be a weighted $\infty$-norm \begin{equation*} \|(v_1,...,v_N)\|_W\equiv \max_{j=1,...,N}\|v_j\|_j, \hspace{2mm} \|v_j\|_j\equiv\sup_{t\in[t_{j-1},t_j]}|v_j(t)|_W,\hspace{2mm} |v_j|_W\equiv |W v_j| \end{equation*} where the weight, $W$, is some invertible matrix. In practice, we will take $W$ to be diagonal and choose the diagonal elements to balance the magnitude of the errors in the different components of the solution. One could consider the generalization where the weight is allowed to vary between subintervals but this will not be explored here. $Y_1$ is a Banach space and the maps that send any $y\in Y_1$ to its value at any $t\in\{0,1\}\cup_{j=1}^N(t_{j-1},t_j)$ or at $t_j^{\pm}$ ($+$ denots limit from above, $-$ limit from below) are well defined bounded linear maps. Elements of $Y_1$ can be identified with elements of $L^\infty$ and integration or differentiation (of piecewise $C^1$) elements of $Y_1$ will be defined using this identification. Mirroring the continuous case, (\ref{X_def}), define a second space \begin{equation}\label{Y2_def} Y_2=\{v\in Y_1: v(0)=0\}\times\mathbb{R}^n. \end{equation} This is a Banach space under the norm $\|(v,b)\|_{W_1,W_2}=\max\{\|v\|_{W_1},|b|_{W_2}\}$ where $W_i$ are any invertible matrices. Finally, note that the $X_i$'s are closed subspaces of the $Y_i$'s. For simplicity we will take $W_1=W_2=W$. This will allow us to drop the subscript $W$'s from the above norms in the following sections. Having specified the functional analytic framework, the definition of nonlinear BVPs can be extended to this setting by defining $G:Y_1\rightarrow Y_2$ by the same formula as in (\ref{G_def}). Note that any zero of this map is automatically continuous, and so solutions of the generalized problem are still in 1-1 correspondence with solutions of the original BVP. The extended operator $G$ is still Fr\'echet-$C^1$, with the same formula for the derivative (\ref{DF_formula}). Note, however, that $D_y f(s,y_0(s))$ is no longer continuous as a function of $s$ for a general $y_0\in Y_1$. Therefore, in the interest of applying Newton-Kantorovich to prove existence, we are lead to consider bounded linear maps $F:Y_1\rightarrow Y_2$ defined as in (\ref{lin_bvp}), where the assumptions on $A(s)$ are relaxed to allow for piecewise continuity on the given mesh. \section{Green's Function for Linear BVPs}\label{sec:greens} We are now in position to discuss the BVP fundamental solution, on which our method will be based. A BVP fundamental solution is defined as a continuous and piecewise $C^1$ solution to \begin{equation*} \Phi^\prime=A\Phi,\hspace{2mm} B_0\Phi(0)+B_1\Phi(1)=I \end{equation*} consisting of invertible matrices. Recall that $A(t)$ is only assumed to be piecewise continuous. Analogously to the IVP fundamental solution, if a BVP fundamental solution exists then it provides an inverse to the BVP operator $F$ through the Green's function, as defined in the following theorem. \begin{theorem}\label{lin_BVP_solution} If a BVP fundamental solution $\Phi$ exists then it is unique and the unique solution to \begin{equation*} v(t)-v(0)-\int_0^t A(s)v(s)ds=r(t),\hspace{2mm} B_0v(0)+B_1v(1)=w \end{equation*} can be written \begin{align*} v(t)=&\Phi(t)(w-B_1r(1))+r(t)+\int_0^1 G(t,s)A(s)r(s)ds \end{align*} where the Green's Function is defined as \begin{align}\label{greens_def} G(t,s)=\begin{cases} \Phi(t)B_0\Phi(0)\Phi^{-1}(s), & \text{if }{s\leq t} \\ -\Phi(t)B_1\Phi(1)\Phi^{-1}(s), & \text{if }{ s>t}. \end{cases} \end{align} Therefore $\Phi$ provides us with an inverse of the operator $F$ from (\ref{lin_bvp}) given by \begin{equation}\label{F_inv_formula} F^{-1}:(r(t),w)\rightarrow\Phi(t)(w-B_1r(1))+r(t)+\int_0^1 G(t,s)A(s)r(s)ds. \end{equation} \end{theorem} Note that the Green's function (\ref{greens_def}) is not to be confused with the notation for the operator associated with a nonlinear BVP (\ref{G_def}). The meaning will be clear from the context. Theorem \ref{lin_BVP_solution} is essentially found in \cite{ascher_petzold}. The only complications beyond the presentation there is that $A(t)$ is only piecewise continuous and $r(t)$ must be taken to be an arbitrary continuous function that vanishes at $t=0$, not necessarily one of the form \begin{equation*} r(t)=\int_0^tq(s)ds \end{equation*} for some continuous $q$, as would be the case when transforming from the differential to the integral form of the BVP. One can directly verify the formula by composing (\ref{F_inv_formula}) with (\ref{lin_bvp}) and vice versa. \section{Proving Invertibility} We now suppose that we are given piecewise $C^1$ maps \begin{equation*} \tilde \Phi:[0,1]\rightarrow\mathbb{R}^{n\times n},\hspace{2mm}\tilde G:[0,1]\times[0,1]\rightarrow\mathbb{R}^{n\times n}, \end{equation*} which are thought of an approximate fundamental BVP solution and approximate Green's function respectively (but without any a priori knowledge that an exact $\Phi$ actually exists) which were obtained by non-rigorous numerical means. More specifically, there are a set of nodes $0=t_0<...<t_N=1$ such that each $\tilde\Phi|_{[t_{i-1},t_i]}$ is $C^1$ and each $\tilde G|_{[t_{i-1},t_i]\times[t_{j-1},t_j]}$ is $C^1$, except when $i=j$, in which case $\tilde G$ also has a discontinuity along the diagonal. We will construct $\tilde G$ using $\tilde\Phi$ and an approximation to its inverse by replacing the corresponding exact quantities in the definition of the Green's function (\ref{greens_def}), but this is not strictly necessary. Our strategy for proving the existence of $F^{-1}:Y_2\rightarrow Y_1$ is to mimic the formula (\ref{F_inv_formula}) for $F^{-1}:X_2\rightarrow X_1$ in terms of the fundamental solution but using the approximate fundamental solution instead. More precisely, replacing $\Phi$ with $\tilde\Phi$ and $G$ with $\tilde G$ in the formula for $F^{-1}$ yields a bounded linear map $H:Y_2\rightarrow Y_1$. $H$ will be used to find conditions under which we can prove the invertiblility of $F$. This will rely on the following well known result concerning perturbations of the identity, see for example \cite{folland}. \begin{lemma}\label{I_perturb_lemma} Let $A:X\rightarrow X$ be a bounded linear map on a Banach space. If $\|I-A\|<1$ then $A$ has a bounded inverse that satisfies \begin{equation*} \|A^{-1}\|\leq\frac{1}{1-\|I-A\|}. \end{equation*} \end{lemma} In fact, a slight generalization of this result will be needed, given below. \begin{lemma}\label{inverse_lemma} Let $A:X\rightarrow Y$, $B:Y\rightarrow X$ be bounded linear maps where $X,Y$ are Banach spaces. If $\|I-AB\|\leq\alpha$ for $\alpha<1$ where $I$ is the identity on $Y$ then $A$ is surjective. If $A$ is also injective then $A$, $B$ are invertible, \begin{equation*} \|A^{-1}\|\leq\frac{\|B\|}{1-\alpha} \text{ and } \|A^{-1}-B\|\leq \frac{\alpha \|B\|}{1-\alpha}. \end{equation*} \end{lemma} \begin{proof} By the above lemma, $AB$ has a bounded inverse with \begin{equation*} \|(AB)^{-1}\|\leq\frac{1}{1-\|I-AB\|}\leq \frac{1}{1-\alpha}. \end{equation*} In particular $A$ is surjective. It is injective by assumption so by the open mapping theorem it has a bounded inverse. This implies $B$ has a bounded inverse as well and \begin{equation*} \|A^{-1}\|=\|BB^{-1}A^{-1}\|=\|B(AB)^{-1}\|\leq \frac{\|B\|}{1-\alpha}, \end{equation*} \begin{equation*} \|A^{-1}-B\|=\|A^{-1}(I-AB)\|\leq \|A^{-1}\|\|I-AB\|\leq \frac{\alpha \|B\|}{1-\alpha}. \end{equation*} \end{proof} \subsection{Applying the Fredholm Alternative}\label{sec:fred_alt} In finite dimensions and when $X$ and $Y$ have the same dimension then the hypothesis of injectivity of $A$ in Lemma \ref{inverse_lemma} is not needed but in infinite dimensions it is required in general. However, when we have the appropriate element of compactness, the situation in infinite dimensions mirrors the finite dimensional case. More specifically, recall one consequence of the Fredholm alternative, see for example \cite{eidelman2004functional}. \begin{theorem} Let $X$ be a Banach space and $K:X\rightarrow X$ be a compact operator. Then $I-K$ is injective iff it is surjective. \end{theorem} If a Fredholm alternative-like condition holds for $F$ then surjectivity will imply injectivity, and hence $\|I-FH\|<1$ will imply the invertiblity of $F$. In addition, this will provide a bound on $F^{-1}$ per Lemma \ref{inverse_lemma}. These are the two ingredients that are needed in order to utilize the Newton-Kantorovich theorem. With slight modifications of the above we could just as well work with $I-HF$ but find this less appealing for a reason we point out in Appendix \ref{app:I-HF}. In order to more easily apply the Fredholm alternative, we reformulate the BVP operator $F$ in terms of a map from a single Banach space to itself. Define $\tilde F:Y_2\rightarrow Y_2$ by \begin{align*} \tilde F[v,w](t)=&\left(v(t)-\int_0^t A(s)(w+v(s))ds,B_0w+B_1(w+v(1))\right)\\ =&(v(t),w)-\left(\int_0^t A(s)(w+v(s))ds,(I-B_0-B_1)w-B_1v(1)\right)\\ \equiv& (I-K)[v,w](t). \end{align*} We have $F[v]=\tilde F[v-v(0),v(0)]$ and so $F$ is surjective or injective if and only if $\tilde F$ is. The second component of $K$ has finite rank and the first is an integral operator with bounded kernel over a bounded domain and whose image consists of continuous functions, hence the compactness of $K$ follows from an application of the Arzela-Ascoli theorem. This implies that the Fredholm alternative applies to $\tilde F$ and hence, showing that $\|I-FH\|\leq\alpha<1$ is sufficient to prove that $F$ has a bounded inverse with \begin{equation}\label{F_inv_bound} \|F^{-1}\|\leq\frac{\|H\|}{1-\alpha},\hspace{2mm} \|F^{-1}-H\|\leq \frac{\alpha \|H\|}{1-\alpha}. \end{equation} We will now discuss how to bound $\|I-FH\|$ and $\|H\|$. \subsection{Bounding $\|I-FH\|$} In Appendix \ref{app:I-FH} a formula for $I-FH$ is derived. For each $t\not\in\{ t_i\}_{i=0}^N$ let $j_t=\max\{j:t_j<t\}$. The two components of $I-FH$ are given by \begin{align}\label{I-FH_components} &(I-FH)[r,w]_1(t)\\ =&-\int_0^t\tilde \Phi^\prime(s)-A(s)\tilde\Phi(s)ds(w-B_1r(1))+\int_0^t(\tilde G(z^-,z)-\tilde G(z^+,z)+I)A(s)r(s)ds\notag\\ &+\sum_{j=1}^{j_t}(\tilde\Phi(t_j^-)-\tilde\Phi(t_{j}^+))(w-B_1r(1))+\int_0^1\left[\sum_{j=1}^{j_t}(\tilde G(t_j^-,z)-\tilde G(t_{j}^+,z))\right]A(z)r(z)dz\notag\notag\\ &-\int_0^t\int_0^1(\partial_s \tilde G(s,z)-A(s) \tilde G(s,z))A(z)r(z)dzds,\notag\\ &(I-FH)[r,w]_2\notag\\ =&[I-B_0\tilde\Phi(0)-B_1\tilde\Phi(1)](w-B_1r(1))-\int_0^1 \left(B_0\tilde G(0,s)+B_1\tilde G(1,s)\right)A(s)r(s)ds\notag \end{align} where superscripts $+$ and $-$ denote the limits from above and below respectively. For completeness, the analogous expression for $I-HF$ is given in Appendix \ref{app:I-HF}. The most straightforward way to use (\ref{I-FH_components}) to bound the norm of $I-FH$ is to define $\tilde G$ using (\ref{greens_def}), replacing the exact quantities with $\tilde\Phi$ and its approximate inverse, and then use the submultiplicative property of induced matrix norms and bound each of the quantities in the Green's function individually. This leads to an estimate whose operation count scales linearly in the number of subintervals, $N$. If $|\Phi|$ and $|\Phi^{-1}|$ are not too large over $[0,1]$ then this is a good strategy. However, this is often not the case, even in situations where the BVP is stable. We illustrate this phenomenon with the same test problem introduced in (\ref{test_prob}). \subsection{Behavior of the Test Problem}\label{sec:test_prob} Consider again the simple test problem, (\ref{test_prob}). The fundamental solution and its inverse are \begin{equation*} \Phi(t)=\frac{1}{\sinh(b)}\left( \begin{array}{cc} \sinh(b-t) & \sinh(t) \\ -\cosh(b-t) & \cosh(t)\\ \end{array} \right),\hspace{2mm} \Phi^{-1}(t)=\left( \begin{array}{cc} \cosh(t) & -\sinh(t) \\ \cosh(b-t) & \sinh(b-t)\\ \end{array} \right). \end{equation*} While $\|\Phi\|$ is bounded, $\|\Phi^{-1}\|$ diverges as $b\rightarrow\infty$. The general notion of BVP stability only requires the norm of the fundamental solution and of the Green's function to be small, but not the inverse of the fundamental solution. See for example \cite{ascher_petzold} for details. Therefore, in general we must abandon any estimate based on bounding $\|\tilde\Phi\|$ and $\|\tilde\Phi^{-1}\|$ separately and use one based on bounding the entire Green's function. The test problem (\ref{test_prob}) does have a bounded Green's function \begin{equation*} G(t,s)=\begin{cases} \frac{1}{\sinh(b)}\left( \begin{array}{cc} \sinh(b-t)\cosh(t)&-\sinh(b-t)\sinh(t) \\ -\cosh(b-t)\cosh(t) & \cosh(b-t)\sinh(t)\\ \end{array} \right), & \text{if }{s\leq t} \\ -\frac{1}{\sinh(b)}\left( \begin{array}{cc} \sinh(t)\cosh(b-t)&\sinh(t)\sinh(b-t) \\ \cosh(t)\cosh(b-t) & \cosh(t)\sinh(b-t)\\ \end{array} \right), & \text{if }{ s>t}. \end{cases} \end{equation*} We emphasize that $\|G\|=\mathcal{O}(1)$ does not arise from a large cancellation of the computed results. Rather, at each endpoint the boundary conditions annihilate the growing modes of $\Phi^{-1}$. This suggests that we should not run into the same degree of numerical cancellation issues using the BVP method as opposed to the IVP method. \section{A Sharper Bound}\label{sec:sharper_bound} In this section we show how to compute a sharper bound on $\|I-FH\|$ by bounding the Green's function as a whole. Bounding the second component of $I-FH$ is straightforward so we focus on the first. To this point the only assumption made about $\tilde\Phi$ and $\tilde G$ is that they are piecewise $C^1$. In practice, a non-rigorous numerical algorithm will provide us with approximate solution values $\tilde\Phi_j$ and $\tilde G_{i,j}$ at the centers of the intervals $[t_{i-1},t_i]$ and rectangles $[t_{i-1},t_i]\times [t_{j-1},t_j]$. For example, in our computations we obtain the $\tilde\Phi_j$ by numerically solving the BVP and, for $i\neq j$, define \begin{equation*} \tilde G_{i,j}=\begin{cases} \tilde\Phi_iB_0\tilde\Phi_1\Psi_j, & \text{if }{j< i} \\ -\tilde\Phi_iB_1\tilde\Phi_N\Psi_j, & \text{if }{ j> i} \end{cases} \end{equation*} where $\Psi_j$ is a numerical approximation to $\Phi_j^{-1}$. On the diagonal $i=j$ two different versions of $\tilde G_{i,i}$ are needed. These are denoted by \begin{equation*} \tilde G^-_{i,i}=\tilde\Phi_iB_0\tilde\Phi_1\Psi_j,\hspace{2mm} \tilde G^+_{i,i}=-\tilde\Phi_iB_1\tilde\Phi_N\Psi_j, \end{equation*} corresponding to the lower and upper halves of the triangle respectively. Given the values at the nodes, $\tilde G|_{[t_{i-1},t_i]\times [t_{j-1},t_j]}$ and $\tilde \Phi|_{[t_{j-1},t_j]}$ are constructed to be the polynomials on each subinterval or rectangle that satisfy the equations \begin{equation*} \partial_t G(t,s)=A(t) G(t,s),\hspace{2mm} \partial_s G(t,s)= -G(t,s)A(s),\hspace{2mm}G(t_{i-1/2},t_{j-1/2})=\tilde G_{i,j} \end{equation*} and \begin{equation*} \Phi^\prime=A\Phi,\hspace{2mm} \Phi(t_{j-1/2})=\tilde\Phi_{j} \end{equation*} up to some order $m-1$. Formulas for the coefficients and the ODE residual are given in Appendix \ref{app:taylor}. Again, on the rectangle with $i=j$ different versions of $\tilde G$ are needed, one corresponding to the upper half of the triangle and one to the lower, constructed from $\tilde G^+_{i,i}$ and $\tilde G^-_{i,i}$ respectively. The resulting $\tilde\Phi$ and $\tilde G$ will be piecewise smooth but not continuous. This is the reason we generalized the solution space in section \ref{sec:gen} to allow for discontinuities. We emphasize that in the absence of jump discontinuities at the mesh points, this method reduces to a piecewise Taylor model IVP method for $\Phi$, see for example \cite{Nedialkov199921}. While well suited for initial value problems, such methods perform extremely poorly for BVPs with an unstable IVP such as the examples in section \ref{sec:num_tests}. Therefore, having accurate $\tilde\Phi_{j}$'s as inputs is crucial to the success of the method; the Taylor series simply ensure sufficiently small error over the interior of each subinterval of the mesh. Using Appendix \ref{app:taylor} we can write $A\tilde\Phi(t)-\tilde\Phi^\prime(t)= R_j(t)\tilde \Phi_{j-1}(t-t_{j-1})^m$ on each $[t_{j-1},t_{j}]$, and similarly for $\tilde G$, and break the integrals in (\ref{I-FH_components}) into sums of integrals over the subintervals of the mesh to obtain \begin{align} &\|(I-FH)_1\|\label{I-FH_bound2_1}\\ \leq&\sum_{j=1}^N\sum_{k=1,j\neq k}^N2h_k\frac{(h_j/2)^{m+1}}{m+1}|\tilde G_{j,k}|\|I+F_k(z)\|_k\|A\|_k\| R_j\|_j\notag\\ &+\sum_{j=1}^{N-1}\sum_{k=1}^Nh_k|(I+E_j(t_j))\tilde G_{j,k}-(I+E_{j+1}(t_j))\tilde G_{j+1,k}|\|I+F_k(z))\|_k \|A\|_k\notag\\ &+\sum_{j=1}^Nh_j\frac{(h_j/2)^{m+1}}{m+1}\left[\|\tilde G^-_{j,j}(I+F_j(z))\|_j+\|\tilde G^+_{j,j}(I+F_j(z))\|_j\right]\|A\|_j\|R_j\|_j\notag\\ &+\sum_{j=1}^Nh_j\|(I+E_j(z))\left(\tilde G^+_{j,j}-\tilde G^-_{j,j}\right)(I+F_j(z))+I\|_j \|A\|_j\notag\\ &+\sum_{j=1}^{N-1}|(I+E_j(t_j))\Psi_j-(I+E_{j+1}(t_j))\Psi_{j+1}|(1+|B_1|)\notag\\ &+\sum_{j=1}^N2\frac{(h_j/2)^{m+1}}{m+1}(1+|B_1|)\|R_j\Psi_j\|_j,\notag \end{align} \begin{align} &|(I-FH)_2|\label{I-FH_bound2_2}\\ \leq&\sum_{j=1}^Nh_j\|(B_0(I+E_1(0))\tilde G_{1,j}+B_1(I+E_N(1))\tilde G_{N,j})(I+F_j(s))\|_j\|A\|_j\notag\\ &+|I-B_0(1+E_1(0))\Psi_{1}-B_1(1+E_N(1))\Psi_{N}|(1+|B_1|),\notag \end{align} where the subscript $j$ on $\|\cdot\|_j$ indicates that the supremum is taken only over $[t_{j-1},t_j]$. $E_j$ and $F_j$ are defined in (\ref{Ej_def}) and (\ref{Fj_def}) respectively. In bounding $R_j$ on $[t_{j-1},t_j]$ a means for analytically bounding the error term in the Taylor series expansion of $A(t)$ is needed. This can be achieved from an interval matrix valued extension for $A(t)$ and its derivatives. To bound the polynomial terms in the $R_j$, $E_j$, and $F_j$'s we use the simple method of summing the bounds on the norms of each monomial. Finally, to use (\ref{F_inv_bound}) to bound the norm of $F^{-1}$ a machine computable bound on $\|H\|$ is needed. \begin{align} &\|H\|\label{H_bound}\\ \leq &\max_l\bigg[1+h_l\|(1+E_l(t))\|_l\max\{\|\tilde G^-_{l,l}(I+F_l(s))\|_l,\|\tilde G^+_{l,l}(I+F_l(s))\|_l\}\|A\|_l\notag\\ &+\|(1+E_l(t))\Psi_l\|_l(1+|B_1|)+\sum_{j=1,j\neq l}^Nh_j\|(1+E_l(t))\|_l|\tilde G_{l,j}|\|I+F_j(s)\|_j\|A\|_j\bigg].\notag \end{align} The leading order term in $N$ of the complexity of computing these quantities by the above formulae is $\mathcal{O}(n^3N^2)$ where $n$ is the dimension of the ODE system. In particular, it does not depend on $m$, and so from the point of view of complexity, the degree of $\tilde\Phi$ can be set as high as is necessary for the Taylor series errors over the subintervals to be negligible, making the boundary conditions and jump terms the main sources of error. Parallelizing the computation of these bounds is straightforward.\\ These results are summarized in the following theorem. \begin{theorem}\label{main_theorem} Let $Y_1$ and $Y_2$ be defined by (\ref{Y1_def}) and (\ref{Y2_def}) respectively and consider the two point BVP $F[v]=(r,w)$ defined by $F:Y_1\rightarrow Y_2$, \begin{equation*} F[v](t)=\left(v(t)-v(0)-\int_0^t A(s)v(s)ds,B_0v(0)+B_1v(1)\right) \end{equation*} where $A(t):[0,1]\rightarrow \mathbb{R}^{n\times n}$ is piecewise $C^m$ on the mesh $\{t_i\}_{i=0}^N$ for some $m\geq 1$ and $B_i\in\mathbb{R}^{n\times n}$. Let $\tilde\Phi$ and $\tilde G$ be constructed as discussed in section \ref{sec:sharper_bound}. Define $H:Y_2\rightarrow Y_1$ by \begin{align*} H[r,w](t)&=\tilde\Psi(t)(w-B_1r(1))+r(t)+\int_0^1 \tilde G(t,s)A(s)r(s)ds. \end{align*} If $\|I-FH\|\leq\alpha <1$, as computed by (\ref{I-FH_bound2_1}) and (\ref{I-FH_bound2_2}), then $F^{-1}:Y_2\rightarrow Y_1$ exists, is bounded, and \begin{equation*} \|F^{-1}\|\leq\frac{\|H\|}{1-\alpha}. \end{equation*} \end{theorem} \section{Validated Solution of Inhomogeneous Linear BVPs}\label{sec:inhomogeneous} For the linear test problems presented in the following section, it will be useful to replace the Newton-Kantorovich theorem with a simpler and more explicit a posteriori error bound. Suppose we have proven $\|I-FH\|\leq\alpha <1$ so that $F$ is invertible. Let $v$ be the exact solution of $F[v]=(r,w)$ and suppose we are given an approximate solution $\tilde v$. Then by (\ref{F_inv_bound}), \begin{equation}\label{inhomo_err} \|v-\tilde v\|=\|F^{-1}[(r,w)-F[\tilde v]]\|\leq \|F^{-1}\|\|F[\tilde v]-(r,w)\|\leq\frac{\|H\|}{1-\alpha}\|F[\tilde v]-(r,w)\|. \end{equation} Suppose $\tilde v$ is piecewise continuous on the mesh $0=t_0<...<t_N=1$ and that $A(t)$ and $r(t)$ are piecewise $C^m$, hence on each subinterval $[t_{j-1},t_j]$ we have Taylor series with remainders \begin{equation*} A(t)=\sum_{l=0}^mA_l (t-t_{j-1})^l+R^A_j(t)(t-t_{j-1})^m,\hspace{2mm} r(t)=\sum_{l=0}^mr_l (t-t_{j-1})^l+R_j^r(t)(t-t_{j-1})^m. \end{equation*} The components of the residual in (\ref{inhomo_err}) can then be bounded. \begin{align} &\|(F[\tilde v]-(r,w))_1\|\leq\notag\\ & \sup_{t\in[t_{j-1},t_j]}|\tilde v(t)-\tilde v(0)-\sum_{k=1}^{j-1}\int_{t_{k-1}}^{t_k} A_k(s)\tilde v(s)ds-\int_{t_{j-1}}^{t} A_j(s)\tilde v(s)ds-r_j(t)|\label{int_line}\\ &+\sum_{k=1}^j \frac{h_k^{m+1}}{m+1}\|R^A_k\|_k\|\tilde v\|_k+h_j^m\|R_j^r\|_j,\notag\\ &|(F[\tilde v]-(r,w))_2|\leq |B_0\tilde v(0)+B_1\tilde v(1)-w|\notag \end{align} where, as above, $\|\cdot\|_j$ denotes the sup norm over $[t_{j-1},t_j]$. Similar to the approximate fundamental solution, given values $\tilde v_j$ at the nodes, $ \tilde v(t)$ is constructed on $[t_{j-1},t_j]$ to be the polynomial \begin{equation*} \tilde v(t)=\sum_{l=0}^m c_l(t-t_{j-1})^l, \hspace{2mm} c_0=\tilde v_{j-1},\hspace{2mm} c_{l+1}=r_{l+1}+\frac{1}{l+1}\!\!\!\!\!\!\!\!\!\!\!\sum_{i=\max\{0,l-m\}}^{l} \!\!\!\!\!\!\!\!\!\!A_i c_{l-i}\text{ for } l=0,...,m-1. \end{equation*} This causes the degree $1$ through $m$ terms in (\ref{int_line}) to vanish. We don't have the freedom to choose the zero order term; it is the analog of the jump terms from (\ref{I-FH_components}). Analytically evaluating the integrals in (\ref{int_line}), the bound on the first component of the residual becomes \begin{align*} &\|(F[v]-(r,w))_1\|\\ &\leq\max_j \bigg[|\tilde v_{j-1}-\tilde v_0-\sum_{k=1}^{j-1}\int_{t_{k-1}}^{t_k}A_k(s)\tilde v(s)ds-r(t_{j-1})|\\ &+\sum_{l=m}^{2m}\frac{h_j^{l+1}}{l+1}\sum_{i=\max\{0,l-m\}}^{\min\{m,l\}}|A_ic_{l-i}|+\sum_{k=1}^j \frac{h_k^{m+1}}{m+1}\|R^A_k\|_k\|\tilde v\|_k+h_j^m\|R_j^r\|_j\bigg],\\ &\text{where }\int_{t_{k-1}}^{t_k}A_k(s)\tilde v(s)ds=\sum_{l=0}^{2m}\frac{h_k^{l+1}}{l+1}\sum_{i=\max\{0,l-m\}}^{\min\{m,l\}}\!\!\!\!\!A_ic_{l-i}. \end{align*} The bound is now in a form that can easily programmed for automated rigorous evaluation using interval arithmetic or estimation in traditional floating point. We also note that the cost is just $\mathcal{O}(N)$ as opposed to the $\mathcal{O}(N^2)$ cost of bounding $\|F^{-1}\|$. \section{Numerical Tests}\label{sec:num_tests} In (\ref{I-FH_bound2_1}) and (\ref{H_bound}) bounds were given on $\|I-FH\|$ and $\|H\|$ that involve finitely representable objects and finite operations with which we can perform validated computations, namely the piecewise polynomials $\tilde\Phi(t)$ and $\tilde G(t,s)$ and the interval valued extension of $A(t)$ and its derivatives. For rigorous bounds, we implemented this using the MATLAB interval analysis package INTLAB \cite{Ru99a} by creating simple routines that perform algebraic operations with polynomials whose coefficients are interval matrices. With such functionality, the code that produces the desired bound on $\|I-FH\|$ is a simple translation of (\ref{I-FH_bound2_1}) and (\ref{I-FH_bound2_2}). In this section, the applicability and accuracy of the BVP method is assessed using a pair of singularly perturbed test problems, taken from \cite{Lee}, as well as a nonlinear example, the existence of a periodic orbit in the Lorenz system. To obtain usable bounds, even the exponentially small parts of the approximate fundamental solution need to be resolved with the same relative accuracy as the larger portions, otherwise the loss of accuracy will prevent the growing modes from being annihilated when forming the approximate Green's function. For problems with quickly decaying modes, the absolute tolerance of the non-rigorous solver must be set near the underflow limit in order to achieve this. This is done in the first two test problems. Tests were performed using a uniform mesh. There is likely room for improvement by intelligently selecting the mesh but we did not employ such methods here. In table \ref{table:tests} the mesh size was chosen by adding points until the error approximately stabilized. \subsection{Example 1: Turning Point} Our first test problem is a singularly perturbed Airy equation \begin{equation*} \epsilon v^{\prime\prime}-(t-1/2) v=0, \hspace{2mm} v(0)=v(1)=1, \end{equation*} which is a model of the quantum mechanical wave function near a turning point in the potential, here at $t=1/2$. The exact solution is a linear combination of Airy functions and is shown in figure \ref{fig:airy_sol}, where $\epsilon=10^{-6}$. The solution exhibits dense oscillations for $t<1/2$ and a boundary layer near $t=1$. We used the BVP method in the cases $\epsilon=10^{-4},10^{-5},10^{-6}$. We used $m=15$ with the aim of making the Taylor series and approximate inverse errors negligible, allowing us to focus on how the error in the initial approximation supplied to the BVP method translates into the bound produced by the method. If we let $\text{err}^i_j$ be the $i$th component of the $j$th jump error from (\ref{int_line}) then the weight used in this test has diagonal elements given by \begin{equation*} w_i\sum_j|\text{err}^i_j|=\text{constant},\hspace{2mm} \max_i w_i=1. \end{equation*} This choice prevents overestimation of the error in certain components when they have widely different scales and makes a large difference in the quality of the bounds, up to several orders of magnitude; it was often the difference between the existence test $\|I-FH\|<1$ passing or failing. Since the jump errors can be computed in $\mathcal{O}(N)$ operations prior to the main $\mathcal{O}(N^2)$ computation, this is an effective and practical way to improve the estimates. The test results are shown in table \ref{table:tests}. The fourth column gives the error in solution values on the mesh that are provided to the BVP method, the fifth column is the bound on $||I-FH||$ computed by the method, and the final column is the computed bound on the absolute value of the error in $v$. The a posteriori BVP method does overestimate the error by several orders of magnitude, but the error bounds are still practical. The IVP fundamental solution method would not be usable in this situation, as the fundamental solution matrix has a norm $|Y(1)|\approx 10^{12}$ for $\epsilon=10^{-4}$. The use of an adaptive weighted norm is crucial for the success of the method and greatly improves the error bounds. Without the weight, the existence test $\|I-FH\|<1$ fails even at $\epsilon=10^{-5}$. The BVP method fails on this problem for $\epsilon=10^{-7}$, as $\Phi$ becomes so ill conditioned that its decaying components cannot be resolved to sufficient relative accuracy due to underflow. This point will reappear in the subsequent example as well; the exponentially decaying modes of $\Phi$ (which tend to be the least `interesting' features) cause the most trouble for the BVP method. In other words, while a large norm for the inverse fundamental solution does not impact BVP stability, over/underflow issues mean that it is still the main limiting factor of the BVP method in its current form. However, the problem is much less severe than ill conditioning in the IVP method. \begin{figure} \centerline{\includegraphics[height=6cm]{airy_sol.eps}} \caption{Exact solution to the turning point problem for $\epsilon=10^{-6}$. For $t<1/2$ the system is oscillatory (left) and it has a boundary layer near $t=1$ (right).}\label{fig:airy_sol} \end{figure} \begin{table} \centering \begin{tabular}{ |c |c |c |c|c|c|} \hline \rule{0pt}{2.3ex} Example & $\epsilon$ & N & input error &$\|I-FH\|$& solution error bound \\ \hline \rule{0pt}{2.5ex} Turning Point &$10^{-4}$& $230$ &$1.4\times 10^{-12}$& $5.2\times 10^{-9}$&$3.1\times 10^{-10}$\\ \hline \rule{0pt}{2.5ex} &$10^{-5}$& $250$ &$5.1\times 10^{-8}$ & $6.1\times 10^{-5}$ & $1.2\times 10^{-4}$\\ \hline \rule{0pt}{2.5ex} &$10^{-6}$ & $600$ & $2.2\times 10^{-6}$& $3.6\times 10^{-3}$&$4.0\times 10^{-4}$\\ \hline \rule{0pt}{2.5ex} Potential Barrier & $10^{-5}$ & $350$ & $6.0\times 10^{-8}$ &$7.0\times 10^{-6}$ & $8.7\times 10^{-4}$\\ \hline \rule{0pt}{2.5ex} & $10^{-6}$ & $600$ & $3.9\times 10^{-8}$ &$1.8\times 10^{-4}$ & $1.3\times 10^{-4}$\\ \hline \end{tabular} \vspace{.3mm} \caption{Test results for the BVP method on the example problems.}\label{table:tests} \end{table} \subsection{Example 2: Potential Well} As a second test, consider a quantum mechanical potential well problem \begin{equation*} \epsilon v^{\prime\prime}+((t-1/2)^2-\omega^2)v=0,\hspace{2mm} v(0)=1,\hspace{2mm} v(1)=2 \end{equation*} where we take $\omega=1/4$. The solution is oscillatory on $[0,1/4]\cup[3/4,1]$ and exponentially decaying as one moves towards the center of $[1/4,3/4]$ as seen in figure \ref{fig:potential} (right pane) for $\epsilon=10^{-6}$. We used $m=15$ and the same formula for $W$ as in the prior example. Again, the method fails once the decaying modes start to underflow, which occurs for $\epsilon=10^{-7}$. \begin{figure} \centerline{\includegraphics[height=6cm]{potential.eps}} \caption{Solution to the potential problem for $\epsilon=10^{-6}$ (right).}\label{fig:potential} \end{figure} For this example we don't have an analytical formula for the exact solution, so the input error reported for in table \ref{table:tests} for this problem is obtained by comparing approximate solution values after refining the mesh. In the first two examples, the supremum norm of the exact solution was relatively consistent between the different parameter values, but that is not true for the current example, where $\|v\|\approx 300$ for $\epsilon=10^{-5}$ and is single digits $\epsilon=10^{-6}$. This explains why the absolute error is so similar for the cases tested despite the difference in $\epsilon$. \subsection{Example 3: Lorenz System} As our final example, we study a non-linear problem, the existence of periodic solutions to the Lorenz system. Scaling the independent variable by the (unknown) period $T$, the BVP is \begin{align*} x^\prime=&T\sigma(y-x),\hspace{2mm} y^\prime=Tx(\rho-z)-y,\hspace{2mm} z^\prime=T(xy-\beta z),\hspace{2mm} T^\prime=0 \end{align*} \begin{equation*} x(0)=x(1),\hspace{2mm} y(0)=y(1),\hspace{2mm} z(0)=z(1). \end{equation*} In order to fix the initial point on the periodic orbit an additional boundary condition is needed, which we take to be \begin{equation}\label{phase_cond} x(0)=y(0), \text{ i.e. }x^\prime(0)=0. \end{equation} Let $f,g:\mathbb{R}^4\rightarrow\mathbb{R}^4$ denote the vector field and boundary conditions respectively and $G$ be the corresponding BVP operator, defined as in (\ref{G_def}). In \cite{Coomes,radii_poly}, computer aided proof techniques were used to show existence of periodic orbits to the Lorenz system. In particular, in~\cite{Coomes} existence was shown for parameter values \begin{equation*} \sigma=10,\hspace{2mm} \beta=8/3,\hspace{2mm} \rho=28 \end{equation*} and initial conditions and period $T$ close to \begin{equation*} x(0)\approx-12.78619,\hspace{2mm}y(0)\approx-19.36419,\hspace{2mm}z(0)\approx24,\hspace{2mm} T\approx1.559. \end{equation*} Note that our initial point on the orbit, defined by (\ref{phase_cond}), differs from the above initial conditions. We will test our method on this same orbit, shown in the $x$-$y$ plane in figure \ref{fig:Lorenz_orbit}.\\ \begin{figure} \centerline{\includegraphics[height=6cm]{Lorenz_orbit.eps}} \caption{Periodic orbit for the Lorenz system, projected onto the $x$-$y$ plane.}\label{fig:Lorenz_orbit} \end{figure} A Taylor series approximation to the solution centered at $t_{i-1/2}$ can be derived \begin{equation*} x(t)=\sum_{j=0}^m x_j (t-t_{i-1/2})^j,\hspace{2mm} y(t)=\sum_{j=0}^m y_j (t-t_{i-1/2})^j,\hspace{2mm} z(t)=\sum_{j=0}^m z_j (t-t_{i-1/2})^j,T), \end{equation*} \begin{align*} x_{r+1}=&\frac{1}{r+1}T\sigma( y_r - x_r),\hspace{2mm}y_{r+1}=\frac{1}{r+1} T(\rho x_r-y_r -\sum_{j=0}^r x_j z_{r-j}),\hspace{2mm}\\ z_{r+1}=&\frac{1}{r+1}T(\sum_{j=0}^rx_j y_{r-j}-\beta z_r), \end{align*} where $x_0$, $y_0$, $z_0$, and $T$ are obtained from numerical approximations to the periodic orbit at the centers of the mesh intervals $[t_{i-1},t_i]$. The Jacobian of the Lorenz system vector field $f$ can be Taylor expanded \begin{align*} &D_yf(t,x,y,z,T)\\ =&\left( \begin{array}{cccc} -T\sigma & T\sigma &0 &\sigma(y_0-x_0)\\ T(\rho-z_0)& -T & -Tx_0 &x_0(\rho-z_0)-y_0\\ Ty_0 & Tx_0 &-T\beta & x_0y_0-\beta z_0\\ 0&0&0&0 \end{array} \right)\\ &+\sum_{j=1}^m\left( \begin{array}{cccc} 0 & 0 &0 &\sigma(y_j-x_j)\\ -Tz_j& 0& -T x_j &\rho x_j-y_j-\sum_{k=0}^{j}x_k z_{j-k}\\ Ty_j & Tx_j &0 & -\beta z_j+\sum_{k=0}^{j}x_ky_{j-k} \\ 0&0&0&0 \end{array} \right)(t-t_{i-1/2})^j\\ &+\left( \begin{array}{cccc} 0 & 0 &0 &0\\ 0& 0& 0&-\sum_{r=m+1}^{2m}(t-t_{i-1/2})^{r}\sum_{j=r-m}^mx_j z_{r-j}\\ 0& 0&0 &\sum_{r=m+1}^{2m}(t-t_{i-1/2})^{r}\sum_{j=r-m}^mx_j y_{r-j}\\ 0&0&0&0 \end{array} \right). \end{align*} Combined with theorem \ref{main_theorem}, this can be used to compute upper bounds on the values of the parameters $\beta$, $K$, $\eta$ that appear in the Newton-Kantorovich theorem, see Appendix \ref{newton_th}. The trivial weight $W=I$ was used in this test. Denote the piecewise polynomial approximate solution by $y_0$, let $F=DG(y_0)$, and $H$ be its approximate inverse as in Theorem \ref{main_theorem}. Using $m=15$, an equally spaced mesh of size $N=35$, and a domain $D=B_\epsilon(y_0)$, $\epsilon\approx 2.3\times 10^{-4}$, our method provides a rigorous existence proof with the following parameters \begin{align*} \|G(y_0)\|\leq& 1.5\times 10^{-10},\hspace{2mm} \|I-FH\|\leq 0.19,\hspace{2mm} \|F^{-1}\|\leq 1.1\times 10^5,\\ K= &8\times 10^{-2},\hspace{2mm} s_0\leq 1.8\times 10^{-5}, \hspace{2mm} s_1\geq 2.1\times 10^{-4}\notag. \end{align*} In other words, a solution exists within a $\infty$-norm ball of radius $s_0$ about the approximate solution and the solution is unique within the ball of radius $s_1$. The computation was done in MATLAB using the interval arithmetic package INTLAB \cite{Ru99a} on a 2GHz Intel Core i7 and took 142 seconds. This includes the time to compute the approximate solution $y_0$ and approximate BVP fundamental solution $\tilde\Phi$. These were computed using the MATLAB routines ode15s and bvp5c with absolute and relative error tolerances of $10^{-9}$. \section{Conclusion} We have presented an a posteriori method for proving existence to and deriving rigorous error bounds for two point boundary value problems on a finite interval, summarized in Theorem \ref{main_theorem}. The method applies to general $n$-dimensional systems without any assumption on the form of the vector field. It uses a (non-rigorous) approximation to the Green's function of the BVP to generate the bounds and can be evaluated using interval arithmetic for mathematically rigorous results or in traditional floating point for faster generation of approximate bounds. Because the method is based on the Green's function of a BVP rather than the fundamental solution to an IVP, the method is applicable to cases where the BVP is stable but the corresponding IVP is unstable. An adaptively chosen weight matrix was used in some of the norms. This improves the quality of the final $L^\infty$ error bounds by several orders of magnitude in many of the tests. In section \ref{sec:num_tests} we tested the BVP method on a pair of singularly perturbed linear problems that exhibit such problematic features as dense oscillation and boundary layers. The method is successful in proving existence and producing reasonable and usable error bounds, even when the relevant perturbation parameter becomes moderately small. We have also successfully applied it to rigorously prove the existence of a periodic orbit in the Lorenz system. In the cases tested, the main limiting factor on the success of the method is the extreme ill conditioning of the approximate BVP fundamental solution $\Phi$ as the parameter that controls the singular perturbation is made smaller. Although $\Phi$ itself has moderate sized norm in the tests, it is close to being singular. The BVP method requires computation of the fundamental solution and its inverse to sufficient {\em relative accuracy}; poorly resolved decaying modes will spoil the method and the method always fails once the underflow limit is reached. This commonly occurs for very stiff problems and so, while the BVP method works for moderately stiff problems as we have shown, it fails in the very stiff limit. This issue could likely be addressed in a brute force manner by employing an interval or floating point arithmetic package that allows for exponent values of substantially larger magnitude than are allowed in standard floating point arithmetic, but we did not employ such a technique here.
\section{Introduction} There is no property of quantum theory which is more important than the positivity coming from its operator formulation in Hilbert space. Born's probability interpretation of quantum theory (QT) depends on it, and in quantum field theory (QFT) it is a direct consequence (without invoking Born definition) of modular localization theory together with the fact that all physical (i.e. finite energy) states are KMS states after restricting them to modular localized subalgebras. Hence in QFT there is a direct relation of quantum causal localization in Hilbert space with statistical ensemble probability (the ensemble of observables localized in a spactime region) with which Einstein had no problems \cite{hol} \cite{Sigma}. Yet gauge theory starts from covariant \textit{pointlike} vectorpotentials which are incompatible with a Hilbert space and rather act in a suitably defined indefinite metric Krein space; it is not the linear structure of Hilbert space but rather the for quantum theory indispensable nonlinear positivity aspects which is violated in gauge theory and only partially recovered in terms of gauge invariance. Whereas for free fields it is trivial to recover the associated pointlike field strengths whose application to the vacuum state generate a Hilbert space, in the presence of interactions the indefinite metric becomes deeply enmeshed with the matter fields (which in zero order are free fields in Hilbert space), so that one needs a rather elaborate operator gauge formalism which requires an extension of the Krein space setting by "ghost operators") in order to be able to extract a subset of local observable fields which act in a Hilbert space. In this gauge formalism important physical states (e.g. charged states) remain outside the formalism. Gauge theory describes the vacuum sector i.e. the Hilbert suspace generated by the gauge invariant observables acting on the vacuum but the gauge-variant field (which includes the charge-carrying matter fields of QED) have no physical meaning. The BRST gauge formalism is a consistent combinatorial perturbative formalism which is (apart from the mentioned vacuum setor) outside the functional analytic operator control of quantum theoy. It is somewhat of a miracle that the extension of the BRST $s -invariance to global objects as scattering process leads to results which are not only consistent with the formalism but also pass many observational tests. In the words of Raymond Stora, the main protagonist of BRst gauge theory "gauge theory is a miracle and one does not understand why, as far as we know it, the formal rules seem to cover observational results of particle physics". Most physicists from the older generation (including myself) knew that quantum gauge theory is a successful "placeholder" for a still unkown perturbative renormalization theory for $s\geq 1.$ Stanley Mandelstam and Bryce DeWitt attempted to go beyond gauge theory at a time when important concepts were still missing. \textit{The lack of Hilbert space positivity is a gaping wound in the QFT description of }$s\geq 1~$\textit{interactions and in particular of }$s=1~ \textit{gauge theory. It is the principle aim of the present paper to overcome this limitations and obtain new results and interpretations which gauge theory misses, as correct description of the charge screening properties of interacting massive vectormesons in particular in couplings with Hermitian fields (the Higgs model).} This is very different from the situation in classical gauge theory where pointlike vectorpotentials have a well-defined conceptual status as classical fields and gauge transformations transform classical vectorpotentials into others in such a way that the observable field strengths remain invariant. Vectorpotentials are useful classical objects even though they do not explicitly appear in Maxwell's equations. In contrast to QFT, Hilbert space positivity and probability have no conceptual counterpart in classical theory, and hence it is not surprising that the alleged quantization parallelism between classical and quantum field theories on which Lagrangian quantization is based fails precisely for s\geq 1~$QFTs as the result of a clash of the pointlike nature of the field description with the Hilbert space positivity of quantum theory. This problem is not limited to massless $s\geq 1~$potentials where it is reflected in the impossibility to obtain pointlike covariant vectorpotentials in the $s=1$ from the covariantization of Wigner's unitary positive energy representations of the Poincar\'{e} group. It \ also shows up in the nonrenormalizabilty of pointlike higher spin interactions; with other words behind pointlike nonrenormalizability of s\geq 1$ interactions there hides a weakening of localization in such a way that the interaction becomes renormalizable if one works with stringlocal fields. The aforementioned gauge formalism is a formal trick to uphold pointlike localization at the expense of sacrificing Hilbert space positivity which at the end of the calculations. So quantum gauge theory comes at a high price in that the gauge invariant part only covers a small part of the full QFT although, as mentioned before, the applixation of the gauge formalism outside the vacuum sector leads to unmerited and (quoting Stora) not understood successes. It turns out that the renormalizable stringlocal matter fields can be used to define extremely singular pointlocal fields whose correlation functions are unbounded in momentum space (infinite $d_{sd}~$in $x$-space); however for massless vectormesons only the stringlocal matter fields survive; in that case the strings are rigid \cite{Bu} and different directions are not unitarily equivalent (Lorentz covariance is spontaneously broken \cite{Froe}) which manifests itself in terms of the perturbative logarithmic infrared divergencies. This limitation affects QED, where the absence of a pointlocal physical electron operator requires to substitute missing spacetime derivations of collision processes by momentum space prescriptions in terms of photon-inclusive cross sections for the scattering of charged particles. In order to improve the conceptual understanding of QFT there are two alternatives: either construct the charged sectors via representation theoretical concepts from the vacuum sector, or find a Hilbert space alternative to the Krein space gauge theory. The first strategy is successful for QFT with mass gaps\footnote The DHR reconstruction of the superselected sectors \cite{Haag} and their amalgamation into a field algebra with an inner symmetry \cite{Do-Ro} for which the fields are not necessarily pointlike but could be semi-infinite stringlike.}, but presents difficult and largely unsolved problems in case of the (gauge-invariant) vacuum sector in zero mass gauge theories \cit {Bu-Ro}. The second strategy is to realize that the clash between localization and the Hilbert space structure can also be resolved by ceding on the side of localization; this causes no conceptual problems since pointlike localization in a Krein space outside a Hilbert space setting is anyhow a fake (if used outside the vacuum sector). It turns out that the conceptual price is rather small compared with what one may have expected considering the fact that Hilbert space positivity is a very strong restriction. It simply consists in allowing fields localized on semi-infinite spacelike strings (in addition to those representing pointlike observables). However this requires the elaboration of a new formalism of renormalized perturbation theory which in case of $s=1$ replaces the gauge setting. The basic ideas and their application to second order perturbation theory, as well as their relevance for the future development of the Standard Model, is the main theme of the present note (see also \cite{stringlocal}, \cite{Sigma}). A systematic study of the problems encountered in the generalization of the Epstein-Glaser causal approach is in preparation \cite{Jens1} \cite{Jens2} whereas some preparatory remarks about the new setting can already be found in \cite{MSY} \cite{Rio} \cite{Bros} \cite{stringlocal}. The recognition of the conceptual origin of the problem is not new. Already Wigner in his 1939 representation theory \cite{Wig} of $m=0$ finite helicity representations of the Poincar\'{e} group knew that covariant vectorpotentials which are associated to $(1/2,1/2)$ representations of the Lorentz-group do not occur in the list of possibilities which arises from the covariantization of the unique ($m=0,\left\vert h\right\vert =1$) Wigner representation. This permits an immediate generalization to $s>1$ tensorpotentials; there is no such problem with interaction-free pointlike massive potentials with short distance dimension $d=s+1;$ they do not allow to take massless pointlike limits (the Proca potential,..) although this problem disappears by passing from potentials to field strengths. Zero mass limits however do exist after converting pointlike potentials into their covariant stringlike siblings which are associated with the same Wigner representation as the field strengths; there is no other way for maintaing the Hilbert space positivity and the standard relation with pointlike field strengths. All these observation on interaction-free potentials are well-known. In the case of couplings involving \textit{massive} $s\geq 1$ fields, the problem between pointlike fields and the Hilbert space positivity is more subtle; in this case \textit{the Hilbert space structure clashes with renormalizability}. The remedy is to convert the pointlike interaction into its stringlike analog and show that the renormalized perturbation theory involving covariant fields localized on semi-infinite spacelike strings is well-defined. The reason why this could work is that stringlocal potentials have $d=1$ independent of spin, which permits to construct renormalizable couplings in the sense of power counting ($d_{int}\leq 4$)~for any spi \footnote In the present note the spin is assumed to be integer so the the potentials are bosonic.} in such a way that the local observables will remain pointlike (possibly pointlike composites of stringlocal fields). The constructions in this note will be limited to $s=1;~$this is not only because in that case the new formalism has its simplest realization, but also since it permits confrontations of new results with observational physics; indeed it is the first contact of ideas coming from local quantum physics (LQP is the algebraic approach to QFT \cite{Haag}) with observational properties of the Standard Model with many new theoretical aspects and new ways of explaining experimental observations. Besides providing a basis for the extension of proofs of structural theorems (spin\&statistics, TCP, the LSZ scattering theory from mass gaps,...) for which the Hilbert space positivity of quantum theory is essential and which therefore is not possible in a gauge setting, the existence of physical matter fields and massive Y-M gluons in the new formulation provides for the first time a QFT setting for \textit{how confinement can be related to the infrared divergences of physical fields} in the massless limit. In all those structural properties, the Hilbert space positivity of the operator-algebraic setting (not taken care of in functional settings) plays an essential role, which explains why, apart from the vacuum sector, they are outside the range of gauge theory. Gauge theory is basically a perturbative combinatorial structure to which operator methods which rely on positivity (spectral representations, functional analysis, operator algebraic methods) cannot be applied. In the existing literature one finds the observation that the so-called axial gauge is consistent with Hilbert space positivity. But what prevented to use this observation as a start of a Hilbert space formulation of $s=1\ interactions is the fact that only the interpretation of this unit vector e\ $as a fluctuating spacelike direction of a field $A_{\mu }(x,e)$ localized on the spacelike line $x+\mathbb{R}_{+}e,$ $e^{2}\equiv e^{\mu }e_{\mu }=-1~$leads to a consistent formulation. In this new setting every field has its independent fluctuating $e$-variable, just as the fluctuating x~$which marks the start of a spacelike half-line. This requires in particular that a Lorentz transformation covariantly changes $e$ (which contradicts the gauge-parameter interpretation). The non-covariant axial gauge setting finally fell into disgrace since it leads to confusing entangled ultraviolet-infrared divergence problems for which no solution was found. It turns out that it is precisely the directional fluctuation property in $e$ which reduces the short distance dimension $d=2$ of the pointlike Proca field $A_{\mu }^{P}$ to $d=1$ of the stringlocal $A_{\mu }(x,e)$ and thus render the interaction fit for renormalization since it lowers the short distance dimension of e.g. massive pointlike QED from d_{int}^{P}=5$ to the power-counting compatible value $d_{int}^{S}=4~$for the stringlocal interaction. \ In fact the stringlocal construction of the S-matrix suggests a round-about definition of higher order pointlike interaction densities whose short distance behavior is precisely that expected from the unlimited increase of the momentum space polynomial degree; but in contrast to the direct pointlike setting, which leads to an ever increasing (with perturbative order) number of coupling parameters, the coupling strengths are those of the interaction-defining first order. Although the off-shell correlations of pointlike fields are expected to have short distance dimensions which increase with the perturbative order, this is not the case for the on-shell scattering amplitudes, as will be explained in these notes. In order to avoid conceptual confusions it is important to point out that the string-localization refers to fields and not to particles. Particles remain Wigner particles, and with the exception of the noncompact localized continuous spin Wigner representation spaces all particle spaces remain compact localizable. The new Hilbert space setting has interesting consequences for the Standard Model. On the one hand, as already mentioned before, the Hilbert space nature of stringlocal massive Y-M fields (massive gluons) instead of pointlike unphysical Y-M fields in Krein space opens the possibility of an understanding of confinement in the limit of vanishing vectormeson mass. By applying resummation techniques to leading infrared logs\footnote Similar to the proof \cite{YFS} of the vanishing of scattering amplitudes of charged particles in the presence of only a finite number of photons (in terms of perturbative resummation techniques).} there are good reasons to expect, that by using the vectormeson mass as a natural covariant infrared regulator parameter, the $m\rightarrow 0$ limit vanishes for all correlations which contain besides an arbitrary number of pointlike composites also a stringlike gluon (or "$e$-unbridged" $q-\bar{q}$ pairs); this is the only known interpretation of the meaning of confinement which explains the non-observability of gluons/quarks and at the same time is consistent with the foundational causal localization principle (in a Hilbert space setting!) of QFT. On the other hand it also turns the "Higgs mechanism" from its head to its feet by realizing that the physical content of the Higgs model is nothing else than the renormalizable interaction of \textit{massive vectormesons coupled to Hermitian} (instead of charged) \textit{fields} $H$. In the correct description there is no symmetry-breaking and \textit{the Mexican hat potential is not put in, but is rather induced} in second order from the renormalizable stringlocal reformulation of a pointlike $gA^{P}A^{P}H$ interaction; hence it is not surprising that the numerical coefficients of the induced potential depend on the ratio of the masses of the two fields. There is no place for couplings of massive vectormesons for symmetry-breaking (which symmetry \footnote Local gauge symmetry is not a symmetry but a formalism which permits to extract physics fom a Krein space decription (although it is in itself unphysical it leads to a physical Hilbert space subalgebra).}) by nonvanishing one-point functions of gauge-variant fields; the conceptual difference between an induced Mexican hat potential and one put into the interaction in order to support the incorrect idea of mass creation through spontaneous symmetry breaking in the massless two-parametric scalar QED cannot be bridged by arguments which are consistent with QED and also not with the principles of QFT where masses of the model-defining elementary fields belong to the input data and only boundstate masses of particles interpolated by composite fields are predictions of the model. The Hermitian model shares with its complex counterpart ("massive QED") the screening of its Maxwell charge. Since there is no particle/antiparticle counting charge, the characteristic property of the Hermitian coupling is the screened charge of the Maxwell-current which is the only conserved current of the abelian Higgs model. A conserved current of a spontaneous broken symmetry leads to a divergent charge (this is really the intrinsic definition of a spontaneously broken symmetry) whose large distance divergence is caused by the presence of a massless Goldstone boson (this is the content of Goldstone's theorem). The stringlocal interaction associated to the pointlike $gA^{P}A^{P}H$ brings a stringlocal selfadjoint scalar $\phi (x,e)$ into the game which (unlike the Higgs field $H$) shares its degrees of freedom with those of the massive vectorpotentials i.e. does not result from an additional coupling; the degrees of freedom are not changed by the presence of the "intrinsic escort" field $\phi $. The appearance of these stringlocal intrinsic escorts is a new phenomenon which results from the implementation of Hilbert space positivity in terms of stringlocal fields for $s\geq 1$ interactions and has no counterpart in pointlocal models. The recent proposal for a Hilbert space based formulation is not the first attempt to avoid the Krein space setting of $s=1$ gauge theory. Already in the 60s DeWitt \cite{DeWitt} and Mandelstam \cite{Man} explored the possibility of circumventing gauge theory by implementing interactions directly in terms of Hilbert space-compatible field strengths. But without the awareness of the short distance singularity-reducing role of directional fluctuations of stringlocal vectorpotentials, a renormalizable theory in Hilbert space cannot be formulated. More important for the present Hilbert space based setting was the existence of a powerful structural theorem for QFTs with compact localizable (pointlike generated) observable subalgebras and a positive energy representation of the Poincar\'{e} group for which the energy momentum spectrum contains a mass gap \cite{Bu-Fr}. Whereas the validity of scattering theory in the presence of a mass gap is hardly surprising, the assertion that the generating algebras of the full theory can be localized in (arbitrarily narrow) spacelike cones is somewhat unexpected. The previous observations about possible clashes between the pointlike localization of tensorpotentials and the Hilbert space positivity, as well as their resolution by working instead with stringloal fields, find their natural explanation in this theorem. For zero mass vectormesons one cannot rely on this theorem; in that case the use of the quantum Gauss theorem \cit {Bu} leads to the stringlocal nature of Maxwell-charge-carrying fields \cit {Bu}. Those charged strings are \textit{rigid}, in particular their direction cannot be changed by Lorentz transformations (spontaneous breakdown of Lorentz symmetry \cite{Froe}) The re-translations of these findings from the algebraic "local quantum physics" (LQP) setting \cite{Haag} into that of (operator algebra-) generating fields\footnote Whereas the core of (causally closed compact) double cones are points, that of (causally closed noncompact) spacelike cones are covariant semi-infinite spacelike linear strings.} (operator-valued distributions) means that in any QFT with a mass gap, which contains a pointlike generated observable subalgebra which generates the vacuum sector, the superselected charged sectors can be described in terms of stringlike covariant fields $\Psi (x,e).~$In this terminology "pointlike" is the special case of stringlike namely $e$-independence. To generate QFTs in terms of fields one does in particular not need fields which are localized on hypersurfaces ("branelike"-fields). The theorem does however not say anything about whether a particular concrete model can be generated by pointlike fields or if stringlocal fields are necessary. Here our perturbative results connect the necessity for using stringlike localization with the breakdown of renormalizability for their pointlike counterparts. In particular $s\geq 1~ interactions have an interaction density with short distance scale dimension $d_{int}>4$ require the use of stringlike localization. The connection between nonrenormalizability and weakening of pointlike localization is, according to my best knowledge, a new result which will have consequences on earlier attempts to relate nonrenormalizability of certain models with modified renormalization group behavior \cite{Reuter}. The interesting question of an analog of the Callan-Symanzik equations for stringlocal interactions will remain outside the scope of the present paper. Gauge theories and the stringlike Hilbert space formulation as well as their relation to each other is the main topic of this note. The next section entitled "kinematical prerequisites" presents the construction of the massive stringlike interaction-free vectorpotential A_{\mu }(x,e)$ in terms of their pointlike Proca siblings $A_{\mu }(x)~$and a scalar escort field $\phi (x,e)$ which are all members of the same stringlocal localization class. Whereas the massive Wilson loop in terms of the stringlocal field is equal to that in terms of the Proca field and hence its $e$-independence is manifest, its zero mass limit is $e$-independent in a topologically more subtle way. This explains the breakdown of Haag duality for multiply connected spacetime regions with interesting relations to the quantum mechanical Aharonov-Bohm effect; in fact this duality breakdown is a generic property of all \textit{massless} $s\geq 1$ potentials \cite{B1} \cite{B2}. The use of the indefinite metric pointlike potentials leads to wrong results, in other words the gauge description already breaks down if one considers global gauge invariant objects as Wilson loops; it remains strictly limited to pointlike generated gauge invariant local observables. In the third section the kinematical preparation is used for the calculation of the string-independent second order S-matrix of massive vectormesons interacting with matter (massive scalar QED and the coupling to Hermitian $H -fields\footnote The letter $H$ stands for both: Hermitian and Higgs.}). For the $H$-coupling the "Mexican hat" potential, which has been imposed for the implementation of the alleged symmetry breaking "Higgs mechanism", emerges instead as a second order "induced potential" of the renormalizable coupling of a massive stringlocal vectormeson to \ $H$ fields; \textit{it is not the breaking of gauge symmetry but rather its upholding} which in massive vectormeson-$H$ couplings induces \textit{a second order Mexican hat potential. }In the new stringlocal Hilbert space setting this arises from the string-independence of the second order S-matix. The concluding remarks present a resum\'{e} as well as an outlook. Here additional remarks about consequences of the stringlocal $s\geq 1~ in~Hilbert space (SLF) can be found and the stringlocal scenario for the expected confinement in the zero mass limit of self-interacting massive vectormesons is presented. \section{Kinematical prerequisites} With an s=1 pointlike Proca fiel \begin{eqnarray} &&A_{\mu }^{P}(x)=\frac{1}{(2\pi )^{3/2}}\int e^{ipx}\sum_{s_{3}}u_{\mu }(p,s_{3})a^{\ast }(p,s_{3})+h.c. \label{pro} \\ &&\left\langle A_{\mu }^{P}(x)~A_{\nu }^{P}(x^{\prime })\right\rangle =\frac 1}{(2\pi )^{3}}\int e^{-ip(x-x^{\prime })}M_{\mu \nu }(p)\frac{d^{3}p}{2p_{0 },~~M_{\mu \nu }(p)=-g_{\mu \mu ^{\prime }}+\frac{p_{\mu }p_{\mu ^{\prime } }{m^{2}} \notag \end{eqnarray one can connect two stringlocal field \begin{eqnarray} A_{\mu }(x,e) &=&\int_{0}^{\infty }F_{\mu \nu }(x+se)e^{\nu }ds,~F_{\mu \nu }(x)=\partial _{\mu }A_{\nu }^{P}(x)-\partial _{\nu }A_{\mu }^{P}(x) \label{def} \\ \phi (x,e) &=&\int_{0}^{\infty }A_{\mu }^{P}(x+se)e^{\mu }ds \notag \end{eqnarray which are linearly relate \begin{equation} A_{\mu }(x,e)=A_{\mu }^{P}(x)+\partial _{\mu }\phi (x,e) \label{proca} \end{equation This relation may also be directly derived in terms of the $u$-intertwiners of the three fields\footnote I thank Jens Mund for pointing out that these relations correspond to linear relation between stringlocal intertwiners which have been introduces in \cit {MSY}} defined in (\ref{def}) for the computation of the intertwiners of the fields defined in (\ref{def}). Since in the algebraic LQP setting fields which are relatively local (i.e. members of the same stringlocal Borchers class \cite{St-Wi}) with respect to each other are considered as different "field-coordinatizations" of the same model (the same physics), this zero order relation is the kinematical prerequisite for its continued validity for the interacting potentials (its implementation is part of the renormalization process). For the perturbative Bogoliubov S-matrix only the free relation will be needed. \textit{At this point one begins to understand why the early attempts of Mandelstam and DeWitt failed. What was missing were two interrelated properties, namely the short distance scale dimension-reducing directional fluctuations of stringlocal potentials and the perception that the Hilbert space positivity requires the presence of a scalar "escort" }$\phi ~$\textit of the vectorpotential. This escort field does not generate new degrees of freedom; together with its "mother potential" }$A_{\mu }(x,e)~$\textit{it is simply the result of lowering the }$d_{sd}=2~$\textit{of the Proca potential }$A^{P}(x)$\textit{\ in order to obtain an interaction density which is below the power-counting limit }$d_{int}\leq 4.~$\textit{We will see that the escort enters the interaction density in an essential way. } A useful reading of this relation (equivalent to the picture of lowering of d^{P}=2$ to $d^{S}=1$ by one unit of $d$ going into the $e$-fluctuations) is to say that the derivative of the $d=1$ scalar stringlocal "escort field" \phi ~$compensates the leading short distance singularity at the price of weakening the localization from point- to stringlike. Each vectormeson potential has its $\phi $-companion which shares the degrees of freedom and the mass. It turns out that this mechanism permits a generalization to arbitrary high integer spin in which case there appear $s$ stringlocal escort $\phi ~$fields with spins between $0$ and $s-1$; the scalar $\phi ~ enters with $s$ derivatives and the tensor indices of the different $\phi $ together with the number of derivative in front always add up to $s$. In all cases the relation breaks down in the massless limit since there are neither pointlike Proca fields nor stringlocal massless $\phi ^{\prime }s$ within the mentioned spin range; in this limit only the stringlocal $A_{\mu }$ or equivalenty the linear combination (\ref{proca}) survives- Our main interest is $s=1.$ In this case the integration of the equation along a closed spacelike circle leads t \begin{eqnarray} \doint A_{\mu }(x,e)dx^{\mu } &=&\doint A_{\mu }^{P}(x)dx^{\mu },~m>0 \label{loop} \\ \doint A_{\mu }(x,e)dx^{\mu } &=&\doint A_{\mu }(x,e^{\prime })dx^{\mu },~~\forall e,e^{\prime },~m=0 \notag \end{eqnarray since the integral along a spatial path of a gradient of $\phi ~$is the difference between the endpoint and the initial point which vanishes for coinciding points. The resulting independence of the Wilson loop operator from the direction $e$ corresponds to the expected gauge invariance of the loop. In the zero mass limit, the difference of two line integral over the$~ same curve but with $e^{\prime }s~$pointing into different directions is simply the difference between the $\phi (x,e)$ and $\phi (x,e^{\prime })$. Although the individual $\phi $ are infrared divergent, their difference at the same point (the initial and final point of the Wilson loop) but with different directions stays infrared finite; hence the difference between two Wilson loops over stringlocal potentials with different string direction vanishes (the second line in \ref{loop}). In this case there remains a topological imprint which the string dependence leaves behind. The result is the breakdown of Haag duality for multiply connected spacetime regions. In such a case the algebra of a causally closed region $\mathcal{A( }),$ $\mathcal{O=O}^{\prime \prime }$ is not the same as the commutant of the algebra $\mathcal{A}(\mathcal{O}^{\prime })$~rather one find \begin{equation} \mathcal{A(O})\varsubsetneq \mathcal{A(O}^{\prime })^{\prime } \label{dual} \end{equation for $\mathcal{O}$ a spacelike torus. The "thickened" Wilson loop is an operator in the right hand algebra which is not in $\mathcal{A(O}).$ The physical picture is that there are operators localized in a spacelike separated intertwining torus which have to penetrate the cylinder subtended from the loop into the $e$-direction somewhere whatever direction of $e$ one chooses. The bigger the helicity, the higher is the genus up to which new violations of Haag duality occur. This effect can also be directly derived from the equal time commutation relations of the$~$field strengths without the use of their stringlike tensorpotentials (for $s=1$ this was shown in the unfortunately unpublished work \cite{LRT}). The important point here is that the proof in terms of stringlocal potentials is simpler whereas the use of the Stokes theorem for pointlike potentials in Krein space is misleading. This is the simplest illustration for the necessity of a Hilbert space setting for potentials and the ensuing SLF setting. It should be clear that the stringlike localization is not the result of a playful fancy of particle theorists who set out to "try something else" and look for its potential physical use later on (which led to String Theory \cite{hol}), but rather the consequence of maintaining the Hilbert space positivity of quantum theory also for $s\geq 1$. To avoid misunderstandings about the aim of the present work, it should be said that the central issue of this paper is not the change of the viewpoint about the nature of the QFT analog of the Aharonov-Bohm effect in the Hilbert space setting (instead of the usual gauge-theoretic pointlike setting in Krein space) but rather the changes which Hilbert space positivity causes in the ongoing research of the Standard model (in particular the Higgs issue and confinement). The simplicity of the above presentation of interaction-free spacetime loops is meant as a pedagogic illustration that even free field properties of \textit{global} gauge invariants come out incorrect; the physical range of gauge theory is strictly limited to gauge invariant local observables. Thinking of the functorial relation between Wigner's positive energy representation spaces for the Poincar\'{e} group and interaction-free quantum fields and the associated local nets of operator algebras, it is interesting to note that the operator algebraic breakdown of Haag duality for multiply connected spacetime regions has a \textit{spatial counterpart in modular localized toroidal subspaces of the Wigner representation space}. The spatial counterpart of the stringlocal vectormeson field is the covariant stringlocal Wigner wave functions which together with its opposite frequency part defines a hyperbolically propagating classical wave functions for a classical stringlocal vectorpotential. Different from the standard use of pointlike classical vectorpotentials, the stringlocal vectorpotential extends to spacelike infinity and thus prevents the formation of compact Cauchy data for potentials. The Wilson loops formed with the correct classical potentials, although being causally separated from a magnetic flux inside the Wilson ring, still "feel" its presence because the flux lines have to penetrate the walls of one of the infinite spatial cylinders which are associated to the different choices of the $e^{\prime }s;$ this is a topological imprint which the stringlocal vectorpotential leaves behind if one forms a classical Wilson loop which looses geometric memory of the $e$ of its vectorpotential. This is the way in which the presence of the magnetic flux is perceived despite the geometric causal separation between the potential in the loop and the magnetic flux inside. The Aharonov-Bohm effect is a quasiclassical relic (quantum mechanical matter in a classical vectorpotential) of this breakdown of Haag duality and admits higher helicity generalization ($m=0,\left\vert h\right\vert ~finite$); it looses its exotic appeal by abandoning the idea of pointlike potentials. The observance of this discrepancy between the naive geometric picture of apparent spacelike separation and the classical limit of the more hidden stringlocal nature of quantum potentials removes all feelings of nonlocal magical aspects of the A-B effect. This provides strong support for the Hilbert space formulation of $s\geq 1$ fields as compared to the gauge theoretic setting. It shows that the often as magic perceived mismatch between the naive geometric view and that coming from causal localization of the A-B effect disappears if one permits the more fundamental QFT the chance to tell its classical counterpart to use the correct description (classical stringlocal fields and their topological implication for loops). This philosophy is opposite to that of "quantization" where the (Euler-Lagrange) description is used as a classical crutch to enter the world of QT. Stringlocal fields in Hilbert space are not solutions of Euler-Lagrange equations and the consistency of gauge theory with QT is limited to the vacuum sector (and even there is appears more a matter of luck than of exigency). It is also interesting to note that the better known "Coulomb gauge" is a Hilbert space description which results from the covariant stringlocal potential by directional $e$averaging within a spatial hyperplan \begin{equation} A_{\mu }^{C}(x)=\int A_{\mu }(x,\vec{e})\frac{d\Omega }{4\pi ,~A_{0}^{C}(x)=-\frac{1}{4\pi }\int d^{3}y\frac{\func{div}\vec{E}(\vec{y},t }{\left\vert \vec{y}-\vec{x}\right\vert },~\vec{A}^{C}(x)=\frac{1}{4\pi \int d^{3}y\frac{rot\vec{H}(\vec{y},t)}{\left\vert \vec{y}-\vec{x \right\vert } \label{coul} \end{equation This exposes its covariance property in terms of the action of the Lorentz group on the time-like vector orthogonal to the hypersurface. In massive $s\geq 1$ theories the clash between Hilbert space and pointlike localization and its resolution through the use of stringlike tensorpotentials is reflected in the fact that behind pointlike nonrenormalizability there looms a weakening of localization; the attempt of a pointlike description leads to singular matter fields with short distance dimension $d=\infty $ (unlimited increase with the perturbative order). Mathematically the formal pointlike fields are singular to such an extend that a smearing with all compact supported spacetime testfunctions is not possible. The Wightman localization property can only be recovered in terms or renormalizable stringlocal fields $\Psi (x,e)$ which can be smeared with all Schwartz testfunctions $f(x,e)$\ with compact supports in $\mathfrak{D}$ $\mathbb{R}^{4}~$tensored with $3$-dim. $de~Sitter^{~}),~$see next section .~ $ Linear relations between high dimensional pointlike fields and their lower dimensional stringlike siblings (which for s=1 reduce to (\ref{proca})) are the key for the conversion of nonrenormalizable pointlike interaction densities into affiliated renormalizable stringlocal interactions. Here are two illustrations: \begin{itemize} \item Scalar massive QED with the first order interaction density (products of operators are always Wick-ordered products) which is to be multiplied by a numerical coupling strengt \begin{eqnarray} L^{P} &=&A_{\mu }^{P}j^{\mu },~j^{\mu }=i\varphi ^{\ast }\overleftrightarrow \partial ^{\mu }}\varphi \label{1} \\ L^{P} &=&L-\partial ^{\mu }V_{\mu },~L=A_{\mu }j^{\mu },~V_{\mu }=\phi j_{\mu } \notag \end{eqnarray here we used (\ref{proca}). Whereas the $L^{P}$ has operator short distance dimension $d=5,$ which is too high in order to be within the renormalizable power counting range of $d\leq 4,$ the stringlocal action has $d=4.~$The incriminated dimension 5 has been absorbed into a derivative term where it can be disposed of in the adiabatic limit; in this way the first order S-matrix of the pointlike $L^{P}~$is the same as that of the stringlike $L.~ In fact one does not have to know the polynomial expression, rather its form together with that of $V_{\mu }~$results from the requirement that L-\partial V$ is independent of $e.$ We may say that, given the field content, the renormalizable first order interaction is \textit{self-induced and (in the cases tested up to now) unique.~}This induction principle extends to all orders (see next section). \item The coupling of a massive vectormeson to a Hermitian field i \begin{eqnarray} L^{P} &=&A^{P}\cdot A^{P}H=L-\partial ^{\mu }V_{\mu } \label{2} \\ L &=&A\cdot (A^{P}H+\phi \partial H)-\frac{m_{H}^{2}}{2}\phi ^{2}H,~~V_{\mu }=A_{\mu }^{P}\phi H+\frac{1}{2}\phi ^{2}\partial ^{v}H \notag \end{eqnarray up to renormalizable $H^{3},H^{4}~$self-interactions whose coupling strengths turn out to be fixed by higher order induction so that at the end there is just one first order coupling $g~$and the two masses of the interaction-defining fields. The latter are usually not counted but for comparison with the Higgs mechanism it is helpful to mention them. \end{itemize} In both cases the leading short distance singularity is "peeled off" from the pointlike scalar $L^{P}$ in terms of a divergence of a vector in analogy to the peeling off in (\ref{proca}) by the gradient of a scalar. The $d=5$ of the pointlike interaction density has been converted into the power-counting-compatible $d=4~$of the stringlike density. The split into a stringlocal $d=4\ $interaction and a divergence is unique up to divergence-free additional contributions to $V_{\mu }.$ \section{Higher order string independence of the S-matrix and induced Mexican hat potentials} If the construction of higher order interaction densities TL(x_{1},e_{1})...L(x_{n},e_{n})$ would not involve the time-ordering of operator-valued distributions, the derivatives in $\partial ^{\mu }V_{\mu }$ could be taken outside the time-ordering and the previous relations (\ref{1 \ref{2}) would have a straightforward $n^{th}$ order extension. The singularities at point- and string-crossings prevent this, and in case of interactions between pointlike fields Epstein and Glaser \cite{Ep} established rules for the inductive construction of time ordered products based on the perturbative implementation of causality with minimal scaling degree (which is not larger than the naive high energy divergence degree of corresponding Feynman integrals). Those couplings for which the minimal scaling degree stay finite independent of the perturbative order are called renormalizable. They are those couplings for which the scaling degrees remains finite and the pointlike fields remain localizable. A simple criterion for renormalizability is $d_{L}\leq 4$ where $d_{L~}$is the short distance scaling degree of $L;~$the fulfillment of this power-counting criterion the prerequisite for renormalizability. What renders nonrenormalizable theories physically worthless (apart from possible phenomenological use) is not so much the unbounded increase for p\rightarrow \infty $ but primarily the growth of the number of coupling parameters associated with an ever increasing number of counterterms. An systematic extension of the E-G method to string-crossings has not yet been published \cite{Jens1}, but fortunately such systematics is not yet needed if, as in the present work, one's aim is to direct attention to the \textit{derivation of new concepts and results} from explicit second order model calculations. The main new message is that requirement of string independence of the S-matrix places new restrictions on renormalizations which lead to the concept of \textit{induced normalization terms} i.e. their couplings are uniquely determined in terms of the basic model-defining first order coupling and the masses and spins of the interaction-defining free fields. This is similar to the gauge theoretic setting where the second order $A_{\mu }A^{\mu }~$term of scalar QED is induced from the first order coupling by gauge symmetry. The fundamental difference is however that in the new setting one does not have to invoke a new symmetry, rather the induction is the consequence of causal localization in a Hilbert space setting (modular localization). The main point here is not that the induced counterterm structure is more or less isomorphic\footnote Most interesting are however additional terms in the SLF setting which have no counterpart in the BST gauge formalism.} to that of the gauge setting, but rather that the Hilbert space positivity has additional physical consequences about the structure of global string-independent quantities which the gauge setting cannot reproduce; the simplest free field illustration is the loss of the above topological property in the used of Wilson loops. In second order the $e$-independence is more subtle since the direct definition of the pointlike $TL(x)^{P}L^{P}(x^{\prime })$ fails as a result of nonrenormalizability. However the requiremen \begin{equation} d_{e}(TLL^{\prime }-\partial ^{\mu }TV_{\mu }L^{\prime })=0, \label{indep1} \end{equation where the bracket is considered as a zero differential form in $e$ (on the 3-dim. unit de Sitter space) on which a differential form operator $d_{e}$ acts, is perfectly meaningful since the short distance degrees of $LL$'~and~ V_{\mu }L^{\prime }$ does not exceed 8. A more symmetric form which simultaneously takes care of $e,e^{\prime }$ i \begin{eqnarray} &&d(TLL^{\prime }-\partial ^{\mu }TV_{\mu }L^{\prime }~-\partial ^{\mu \prime }TLV_{\mu }^{\prime }+\partial ^{\mu }\partial ^{\nu \prime T}V_{\mu }V_{\nu }^{\prime })=0,~~d=d_{e}+d_{e^{\prime }} \label{indep2} \\ &&T(LL^{\prime })^{P}:=TLL^{\prime }-\partial ^{\mu }V_{\mu }L^{\prime }~-\partial ^{\mu \prime }LV_{\mu }^{\prime }+\partial ^{\mu }\partial ^{\nu \prime }V_{\mu }V_{\nu }^{\prime } \notag \end{eqnarray Although a direct perturbative treatment of the pointlike second order interaction density would conflict with renormalizability, the indirect definition in terms of the string-independent bracket as in the second line is perfectly reasonable. It is again the encoding of the highest scale dimensions into derivatives of lower dimensional operators within the power counting limit ($d=4n$ in $n^{th}$ order) which permits to define pointlike products whose direct $n^{th}$ order scale dimension $d=4n+n$ would exceed the power-counting limit and, if treated in the usual pointlike setting, would lead to an ever increasing number of coupling parameters since the induction mechanism would be absent. The relations (\ref{indep1},\ref{indep2 ) have straightforward $n^{th}$ order extensions, but here their perturbative model implementation will be limited \ to second order. The similarity of (\ref{proca}) with a gauge transformation suggest that the connection of the stringlocal matter field and its strongly singular (not Wightman-like, pointlike nonrenormalizable) pointlike sibling should be \begin{equation} \psi (x)=e^{-ig\phi (x,e)}\psi (x,e) \end{equation However the conceptual content of these relations is quite different than that of gauge transformations; they do not stand for a symmetry transformation ("gauge symmetry") but rather relate two "field-coordinatizations" in Hilbert space which belong to the same localization class (relative locality with respect to each other). The proof of these off-shell relations in the presence of interactions requires an extension of the St\"{u}ckelberg-Bogoliubov-Epstein-Glaser (SBEG) on-shell formalism. In fact the representation for the S-matrix is a special case of that for fields and their correlation functions; all these formulas require to take the adiabatic limit\footnote The SBEG field construction is opposite to that of LSZ scattering theory; in the latter case on starts from the field correlations and computes the S-matrix as the "cream on the cake".}. Fortunately in massive QED this is not needed since the matter field only enters the interaction in form of the conserved current and the SBEG formalism for the on-shell S-matrix only deals with time-ordered products of free. The expected off-shell connections between renormalizable stringlike fields and their formally nonrenormalizable pointlike siblings contains however the interesting message that the kind of nonrenormalizability addressed in the present work is a result of the clash of enforced pointlike localization with the Hilbert space positivity; a clash which can be removed by passing to the renormalizable stringlocal formulation. The application of (\ref{indep1}) to the model of massive scalar QED (\ref{1 ) requires to expand the time ordered product into Wick-products. The component without contractions fulfills this relation trivially as a consequence of $d_{e}(L-\partial ^{\mu }V_{\mu })=d_{e}L^{P}=0.$ For the 1-contaction component this is not the case since the anomaly $\mathfrak{A}$ \begin{eqnarray} A &:&=d_{e}(T_{0}LL^{\prime }-\partial ^{\mu }T_{0}V_{\mu }L^{\prime })_{1-contr.}=-d_{e}N+\partial ^{\mu }N_{\mu } \\ A^{\prime } &=&A(x,e\longleftrightarrow x^{\prime },e^{\prime }),~~\mathfrak A}=A+A^{\prime } \notag \end{eqnarray where $T_{0}~$refers~to the "kinematical" time ordering for which all derivatives act outside the $T_{0~}~$e.g \begin{eqnarray} \left\langle T_{0}\partial _{\mu }\varphi (x)\partial _{v}^{\prime }\varphi ^{\ast }(x^{\prime })\right\rangle &=&\partial _{\mu }\partial _{v}^{\prime }\left\langle T_{0}\varphi (x)\varphi ^{\ast }(x^{\prime })\right\rangle \notag \\ \left\langle T\partial _{\mu }\varphi (x)\partial _{v}^{\prime }\varphi ^{\ast }(x^{\prime })\right\rangle &:&=\left\langle T_{0}\partial _{\mu }\varphi (x)\partial _{v}^{\prime }\varphi ^{\ast }(x^{\prime })\right\rangle +ig_{\mu \nu }c\delta (x-x^{\prime }) \label{4} \end{eqnarray According to the Epstein-Glaser normalization rules the scaling degree 4 (logarithmically diverging) propagator of derivative of scalar fields permit a scaling degree 4 preserving renormalization in terms of a delta term with a yet undetermined parameter $c$. The contributions to the anomaly coming from the action of the divergence $\partial _{\mu }~$on contractions in T_{0}j_{\mu }(x)j_{\nu }(x^{\prime })|_{1-contr}.$are$~ \begin{equation} \partial ^{\mu }\left\langle T_{0}\partial _{\mu }\varphi (x)\varphi ^{\ast }(x^{\prime })\right\rangle =(\partial ^{\mu }\partial _{\mu }+m^{2})\left\langle T_{0}\varphi (x)\varphi ^{\ast }(x^{\prime })\right\rangle -m^{2}\left\langle T_{0}\varphi (x)\varphi ^{\ast }(x^{\prime })\right\rangle =-i\delta (x-x^{\prime })-reg. \end{equation} It is precisely these numerical anomalies of time-ordered propagators which determine the above operator anomalies. The result of the contraction combinatorics is \begin{equation} N=\varphi ^{\ast }\varphi A\cdot A^{\prime },~N_{\mu }=\delta \varphi ^{\ast }\varphi \phi A_{\mu }^{\prime } \end{equation where the $A_{\mu }^{\prime }$ stands for $A_{\mu }(x,e^{\prime })$ and \delta $ is $\delta (x-x^{\prime }).~$We also use the fact that d_{e}\partial _{\mu }\phi =d_{e}A_{\mu }.~$For the renormalization of T_{0}LL^{\prime }$ the $N_{\mu }~$(which corresponds to the renormalization of $T_{0}V_{\mu }L^{\prime }$) is not important. Hence the relevant $N~$part of the full anomaly is symmetric in $e$ and $e^{\prime } \begin{equation} \varphi ^{\ast }\varphi (A\cdot A^{\prime }+A^{\prime }\cdot A) \end{equation The reader immediately recognizes that this induced contact term which must be added to $T_{0}LL^{\prime }$ corresponds to the quadratic term in the gauge theoretic formulation of scalar QED. In the present setting it is simply the consequence of the string-independence of the S-matrix. The philosophy underlying the present setting suggests to absorb this term into a re-definition (renormalization) of the time ordering by \begin{eqnarray*} T &=&T_{0}~for~propagators~of\text{ }scaling~degree\text{ }d<4 \\ &&T\text{ }as~in~(\ref{4})\text{ }with~c=-1~for\text{ }d=4 \end{eqnarray* In this case the $T_{0}~$contraction of the propagator of the derivative of the fields coming from the anomaly an be absorbed in the redefinition (\re {4}) $T_{0}\rightarrow T$ of the contraction in $T_{0}LL^{\prime }$ so that the $e$-independence (\ref{indep1}) is fulfilled with the $T$ ordering instead of the $T_{0}.$ One can then directly check that (\ref{indep2}) really defines a second order pointlike interaction density without having to introduce a counterterm with a new coupling (as it would be necessary in in the pointlike formalism in Krein space before imposing the gauge invariance condition). In fact the formula (\ref{indep2}) guaranties the independence of the scattering amplitude on the string directions. The string-independence in the Hilbert space setting $d_{e}S=0~$corresponds to $sS=0~$with a nilpotent $s$ in the BRST gauge setting, where the definition of $s~$requires to enlarge the already unphysical Krein space in terms of ghost operators. The computation is analogous, except that there are no$~A_{\mu }^{\prime }.~$The Hilbert space positivity for massive QED leads to on-shell results for $2\rightarrow 2~$scattering which have the same formal appearance as those coming from gauge theory, even though the concepts and calculations are different. In the following we sketch the operator gauge derivation a la Scharf; this should also be seen as a recompensation of the early results of the University of Z\"{u}rich group \cite{Aste} which were left on the wayside by the Standard Model caravan and eventually succumbed to the maelstrom of time\footnote We (Jens Mund and myself) only became aware of this contribution after our work on stringlocal fields \cite{MSY} \cite{Rio}. Here these results have been rewritten into our formalism in order to facilitate comparisons.}. The self-induced $e$-independent $d_{e}(L-\partial V)=0~L,V$ pair of a trilinear $A-H$ coupling (\ref{2}) has its BRST gauge-theoretic counterpart in $s(L^{K}-\partial V^{K})=0$ where the fields now act in the Wigner-Fock Krein space. Apart from the fact that in addition to $u^{K}=s\phi ^{K}~$we also have an anti-ghost field\footnote The ghost field corresponds to $u:=d_{e}\phi ,$ but the anti-ghost has no counterpart since $s\tilde{u}=(\partial A^{K}+m^{2}\phi ^{K})$ and the corresponding SLF expression vanishes since it is a an operator relation in Hilbert space.} $\tilde{u}$\ and the first order induction of a trilinear interaction leads to (we surprise the superscript $K$ for the individual fields) and differences in the normalization of the $\phi ^{K}$ field, the formal expression for the first order gauge interaction is analog to (\ref{2 ) \begin{align} L^{K}& =m\left( A\cdot AH-H\overleftrightarrow{\partial }\phi \cdot A-\frac m_{H}^{2}}{2m^{2}}H\phi ^{2}+bH^{3}+u\tilde{u}H\right) \\ Q_{\mu }^{K}& =m(uA_{\mu }H-\frac{1}{m}u\phi \overleftrightarrow{\partial _{\mu }H)=sV_{\mu }^{K} \notag \\ sA& =\partial u,~sH=0,~s\phi =u,~s\tilde{u}=-(\partial A+m^{2}\phi )\simeq 0 \end{align Up to the $H$ self-interactions these trilinear terms are the unique (up to V_{\mu }~$contributions with vanishing divergence) renormalizable induced first order terms. The third line is the action of $s$ on the individual free fields which act in the Krein space analog of the Fock space. The second order anomalies $sT_{0}LL^{\prime }-\partial ^{\mu }T_{0}Q_{\mu }L^{\prime }~$and the corresponding term with $Q\rightarrow L,~L^{\prime }\rightarrow Q^{\prime }~$(prime denotes $x\rightarrow x^{\prime }$) lead to the additional (induced) contribution \cite{Scharf} \begin{eqnarray} &&T_{0}L^{K}L^{K\prime }+i\delta (x-x^{\prime })(A\cdot AH^{2}+A\cdot A\phi ^{2})-i\delta (x-x^{\prime })R_{scharf} \label{second} \\ &&R_{Scharf}=-\frac{m_{H}^{2}}{2m^{2}}(\phi ^{2}+H^{2})^{2},~V_{Scharf}\equiv g^{2}R+first~order~H,\phi -terms \notag \\ &&V_{Scharf}=g^{2}\frac{m_{H}^{2}}{8m^{2}}(H^{2}+\phi ^{2}+\frac{2m}{g}\phi )^{2}-\frac{m_{H}^{2}}{2}H^{2} \label{pot} \end{eqnarray The remaining step consists in absorbing the quadratic in $A$ contributions into a redefinition of the time-ordered product $T_{0}\rightarrow T~~$(as in the previous case of scalar massive QED) and to verify that the absence of third order anomalies requires to introduce a $cH^{4}$ self-coupling and fixes the parameters $b$ and $c.~$ The net result\ is the $H$-$\phi ~$local potential $R$ of degree 4~(\ref{pot ). The appearance of $g^{-1}~$terms results from writing the potential into the symmetry-breaking form of the Higgs mechanism from where one may read off the Mexican hat parametrization of the Higgs mechanism (the quartic self-coupling of scalar QED and the shift in field space). It shows that that the latter is incompatible with the logic of renormalized perturbation theory, whereas the mechanism of induction potential is its logical second order result. The lesson to be learned is that, whereas at the frontiers of the foundational and sometimes highly speculative research as particle theory one sometimes is led to use metaphors as placeholders for incompletely understood issues, the problem starts when the placeholder nature is not recognized in due time. Admittedly this was not easy since the coupling of Hermitian fields to massive vectormesons which vanish in the massless limit is somewhat unaccustomed for physicists coming from QED. For the details of the anomaly calculations and the derivation of the (up to second order) induced potential (\ref{pol}) we refer to Scharf (formula 4.1.38 \cite{Scharf}). As in the case of the Schwinger-Swieca screening of the Maxwell current of interacting massive vectormesons, we recommend to the readers to take their time to look up the cited historical papers in order to convince themselves how close some individuals came to the correct understanding of interactions involving massive vectormesons. This should have caused the ringing of bells, but the Standard Model Theory community was already in the grip of Big Science; it was too late for a perception and a critical discussion of these findings; more recent attempts to remind the particle theory community of some of these forgotten (or never noticed) results remained without avail \cite{Garcia}. What was still missing in order to realize that the unfortunate broken symmetry picture is inconsistent with the computed result was to notice that the Schwinger-Swieca screening \cite{Sw} of the Maxwell charge of interacting massive vectormesons (the only conserved current in the $H$-model) is completely different from the charge associated to the conserved current of a spontaneously broken symmetry. A deeper sociological analysis of these strange occurrences in the heart of particle theory will be left to historians and archeologists of science. The new SLF Hilbert space formalism does not only confirm these pre-electronic insights\footnote Setting $e=e^{\prime }~$in the scattering amplitudes, the results formally agree (up to differences in the $\phi ^{K}~$versus $\phi $ normalizations).} but also adds an interesting twist to it; the Hilbert space positivity leads to the presence of additional induced contributions to the Mexican hat potentials which however vanish on the diagonal $e=e^{\prime }.~$The definition of pointlike interaction densities \ref{indep2} within the Hilbert space setting also reveals a new property which has no counterpart in the gauge setting. These pointlike densities are special cases of the polynomial expression which one would obtain in second order direct pointlike nonrenormalizable expression, except that instead of new counterterm coupling parameters the latter are now not independent but rather fixed in terms of the first order couplings and masses. As a consequence the on-shell polynomial degree is lower than its off-shell counterpart; with other words the scattering amplitudes have a much better high-energy behavior than that of pointlike correlation functions. This is the momentum space counterpart of "peeling off" high degree derivative terms and dispose them in the adiabatic limit, a mechanism which is difficult to keep track of in momentum space and which shows that the consequences of the Hilbert space positivity for $s\geq 1~$(which requires the use of stringlike localization)$~$ cannot be encoded into Feynman graphs. These deviations of the Hilbert space setting from gauge theory are interesting, but their detailed derivation go beyond the task set for the present paper and will be addressed in separate work \cite{M-S1}. As mentioned in the introduction, the physical credibility of the gauge setting is restricted to the gauge invariant vacuum sector of local observables which strictly speaking does not include the global S-matrix. The frequent formal similarity between the gauge setting and the Hilbert space formulation may be interpreted as an unexpected extension of the physical limitations of gauge theory. In passing it is interesting to see that the screening of the identically conserved Maxwell current can already be seen in zero order of the Maxwell current \begin{eqnarray} j_{\mu }^{M} &:&=\partial ^{\nu }F_{\mu \nu },~~~F_{\mu \nu }=\partial _{\mu }A_{\nu }^{P}-\partial _{\nu }A_{\mu }^{P}=\partial _{\mu }A_{\nu }-\partial _{\nu }A_{\mu } \\ &&in~zero\text{ }order~~~\partial ^{\nu }F_{\mu \nu }\simeq m^{2}A_{\nu },~\partial ^{\mu }A_{\mu }^{P}=0 \notag \end{eqnarray The corresponding zero order Maxwell charge vanishes \begin{eqnarray} &&observation:~Q^{M}=\int j_{0}^{M}(x)d^{3}x,~in~zero\text{ order~j_{0}^{M}(x)\simeq A_{\mu }^{P}(x),~\int A_{0}^{P}(x)=0 \\ &&Thm:The~Maxwell~charge~of\text{ }a~massive~vectormeson~is\text{ }alway \text{ }screened \notag \end{eqnarray Whereas for couplings to Hermitian fields this is the is only conserved current, the complex matter fields of massive QED also admit another conserved current whose conserved charge counts the number of charges minus anti-charges. The charge screening is lost and both currents coalesce in the massless limit. It was Schwinger \cite{Schwinger} who conjectured this property of massive vectormesons which was later proven as a structural (nonperturbative) theorem by Swieca \cite{Sw} \cite{Swieca}. In this context it may be interesting to mention that the conserved current in the case of a spontaneously broken symmetry leads to a divergent charge whose large distance divergence is due to the presence of a zero mass Goldstone boson. Interestingly the structural proof of the Goldstone theorem (which has no relation with nonvanishing one-point functions of fields\footnote The shift in field space is a mnemonic trick to produce a first order interation with such a divergent charge associated to a conserved Goldstore current, but it has no intrinsic physical meaning.}) is also due to Swieca \cite{E-S} \cite{Car}, in fact he emphasized that the large-distance diverging charge is the definition of the meaning of spontaneous symmetry breaking. He probably had the most profound knowledge of both phenomena and he tried to attract attention away from the incorrect Higgs symmetry breaking by using in his publication the terminology "Schwinger-Higgs screening" in his publications. But unfortunately he was unable to stem the growing tide in favor of symmetry breaking. This and the 40 year reign of the Higgs spontaneous breaking mechanism instead says a lot about the present state of post Standard Model. In retrospect it appears very conspicuous that the same arguments and equations which started from massless scalar QED and postulated a symmetry breaking (gauge symmetry?) by performing a "field shift" of a gauge variant field were almost simultaneously independently presented by at least 3 other authors/groups of authors in addition to Peter Higgs. For fairness it should be mentioned that during the first years after these proposals there were also publications which pointed out that gauge symmetry is not a physical symmetry but rather a method to extract local observables from a Krein space setting and as such it is not fit for being broken. Note that the above induced Mexican hat potential can also be obtained by imposing BRST gauge invariance on a massive vectormeson-$H$ interactions \cite{Scharf}; Hence it is it imposition of gauge invariance in interactions of massive $A_{\mu }$ with Hermitian fields and not its breaking in massless scalar QED which is consistent with the foundational principles of QFT. These remarks should not be understood as coming from someone whose research was motivated by the desire to revolutionize particle theory. To the contrary it is the increasing worry that a particle theory which lost its contact with its own pre-electronic past has moved into a blind allay which motivated the author to write this article. An important observation in the aftermath of the discovery of the Standard Model as the "Schwinger-Higgs-Swieca screening mechanism" of massive vectormesons should never have been allowed to be lost in the maelstrom of time. More conceptual and historical observations which receive their support from the ongoing impact of the new SLF Hilbert space setting can be found in \cite{hol}. The many updates of this article since 2011 reflect the continuous development of these ideas. The biggest gain of insight from the new $s\geq 1$ Hilbert space setting\ is expected in the area of self-interactions between massive vectormeson (Y-M couplings). One expects of a setting whose principle task is to classify renormalizable interactions within the Hilbert space setting of renormalizable $d_{e}~$-induced first order interaction densities that it achieves something equivalent to what in classical gauge field theory is obtained with the help of the mathematical fibre-bundle setting. For self-interacting massive vectormesons of \textit{equal masses} the couplings should be restricted to just one coupling strength. In particular for three mutually interacting massive gluons of equal mass the result should look like \begin{eqnarray} &&L=\sum \varepsilon _{abc}(F^{a,\mu \nu }B_{\mu }^{b}A_{\nu }^{c}+m^{2}B^{a,\mu }A_{\mu }^{b}\phi ^{c}),~V^{\mu }=\sum \varepsilon _{abc}F^{a,\nu \mu }(A_{\mu }^{b}+B_{\nu }^{b})\phi ^{c} \label{ren} \\ &&d_{e}(L-\partial ^{\mu }V_{\mu })=0.~Ansatz~L=sum~of~4~terms~in~A,\phi \notag \\ &&\sum f_{abc}^{1}F^{a,\mu \nu }A_{\mu }^{b}A_{\nu }^{c},~\sum f_{abc}^{2}A^{a,\mu }A_{\mu }^{b}\phi ^{c},~\sum f_{abc}^{3}A^{a,\mu }\partial _{\mu }\phi ^{b}\phi ^{c},~\sum f_{abc}^{4}\phi ^{a}\phi ^{b}\phi ^{c} \end{eqnarray where for notational convenience we used the notation $B=A^{P~}$for the pointlike Proca potentials. The last line denotes the 4 structures which can contribute to $L.~$The requirement that $L-\partial V~$is $e$-independent (pointlike) is very restrictive and leads to the first line where the V_{\mu }~$is determined up to a term with vanishing divergence\footnote Similar conclusions follow from the imposition of $s$-invariance of the S-matrix \cite{Scharf}..}. If one defines the pointlike $L^{P}$ as the content of the bracket than the first order pointlike S-matrix is identical to its stringlike counterpart or equivalently: the two $L^{\prime }s~$are adiabatically equivalent (the boundary term from the divergence of $V_{\mu }$ vanishes in the adiabatic limit) \begin{equation} \int L^{P}=\int L,~~~L^{P}\overset{AE}{\simeq }L \end{equation This is the beginning of an extremely restrictive \textit{induction mechanis } which has no counterpart in the nonrenormalizable pointlike $s\geq 1$ setting. For the full Lie-algebra structure one has to proceed to the induced second order \cite{M-S2}.\ These observations generalize those which were already made in the abelian case namely the locality principle \ together with Hilbert space positivity leads to restrictions between couplings which correspond to those of classical gauge theory (the geometry of fibre bundles). Here they are simply the result of the Hilbert space positivity which for interactions which couple $s\geq 1$ fields requires the use of string-localization. There is absolutely no need for any support from the fibre-bundle setting of classical gauge theory; QFT does not need any "crutches" from classical field theory such as those provided by the classical-quantal parallelism of quantization. Any quantum fields obtained from covariantizing Wigner's classification of positive energy representation of $\mathcal{P}$ can be coupled to a scalar density which defines the first order interaction density of a QFT and in case its short distance dimension falls within the power-counting range $d_{sd}\leq 4~$the interaction density is on the best way to define a renormalizable model of QFT. The above "self-induction" mechanism also works for unequal masses; in this case the $f^{\prime }s$ depend also on mass-ratios. The full calculation up to second order will be contained in \cite{M-S2}. Our main point here is that the specification of the prescribed \textit field content together with the }$d_{e}$\textit{\ induction requirement and the renormalizability restriction }$d_{int}\leq 4~$\textit{from power-counting determines the interaction }$L$\textit{\ and the (in the S-matrix adiabatic limit disappearing) divergence contribution of }$V_{\mu } \textit{\ }(again up to $\partial \tilde{V}=0$ terms).\textit{\ Possibly missing restrictions are expected to come from second order. }This means in particular that one does not have the freedom of classical field theory to add interactions; the field content, the power-counting restriction together with the induction mechanism fix the form of the $s\geq 1$ quantum interaction. In particular the second order contributions required by classical gauge invariance cannot be imposed; rather the induction property, behind which hides the powerful Hilbert space positivity, governs the form of the quantum interaction. This is particularly interesting for $s=2$ interactions where the massive stringlocal potentials $g_{\mu \nu }(x,e)$\textit{\ }associated to their pointlike "Proca" siblings $g_{\mu \nu }^{P}(x)~$and two escort fields $\phi $ and $\phi _{\mu ~\ }$already enter the lowest order interaction density; this rigidity of the induction remains in the massless limit and stands in an interesting contrast with the Lagrangian quantization of the classical. Einstein-Hilbert action from which !quantum gravity" could profit. In this connection it is important to point out that renormalization group properties (analogs of the Callen-Symanzik equations) can only be expected to hold for the stringlocal Wightman fields (and not for the singular pointlike fields). Even if the asymptotic freedom property would not have been based on a consistency property but rather on the beta function of an established Callan-Symanzik equation of massive (infrared-finite) QCD in the pointlike gauge setting, its physical relevance would still leave doubts. Only a derivation based on the SLF Hilbert space setting can definitely close this issue. The formal $m\rightarrow 0$ limit in the stringlocal interaction density is generally not sufficient for understanding the content of the massless limit. If there are off-shell (logarithmic) infrared divergences this should be taken as an indication of a radical change in the field-particle relation. Confinement manifests itself in the sense of vanishing of correlations for $m\rightarrow 0~$which contain in addition to pointlike composite also stringlike gluons or quarks. This is expected to be obtained by resummation techniques of the leading $m\rightarrow 0$ logarithmically divergent terms \cite{hol}. \section{Resum\'{e} and outlook} The Hilbert space positivity for higher spin interactions leads to a weakening of localization and requires the replacement of nonrenormalizability of pointlike fields by stringlocal interaction densities. This replaces the BRST gauge theory in Krein space whose physical range of validity of gauge theory is limited to (pointlike generated) gauge invariant local observables; global gauge invariant operators as Wilson loops are already outside its range since they miss the topological origin of the breakdown of Haag Duality (for $s=1$ the Aharonov-Bohm effect).. In the SLF Hilbert space setting fields are generically stringlocal and physical; the subset of pointlike fields correspond to the gauge invariant local observables of gauge theory. The SLF formalism leads to profound conceptual and computational changes. There are 3 basic classes of renormalizable couplings of massive vectormesons: couplings with complex matter, with Hermitian $H~$matter and couplings among themselves including a combination of these 3 classes of couplings; in all cases the masses of vectormesons and the matter fields (in accordance with the principles of QFT) belong to the the model defining field content in terms of which the first order interaction density is defined. The perturbative renormalization formalism secures that these input masses remain equal to the masses of the elementary particles of the model. Up to this point there is no difference to the $s<1$ pointlike renormalization theory. What is however new for $s\geq 1$ is that in a Hilbert space setting renormalizability requires the vectormeson fields to be stringlocal and escorted by fields $\phi $, a property which through higher order perturbation also "infects" the matter fields. As a consequence the demand that certain physical global objects as the S-matrix stay string-independent ($e-$independent) leads to a new phenomenon called induction; instead of counterterms with free coupling parameters the coupling strengths of the induced terms are determined in terms of the interaction-defining first order coupling including the masses of the coupled fields. For the $H-$coupling the induced potential in $H$ and the escort field $\phi $ has the form of a Mexican hat. In the operator gauge setting\footnote In the more common functional setting of perturbation theory (Feynman graphs) the induction mechanism is easily overlooked.} the result is (up to difference in normalization) the same but the SLF derivation of the Mexican potential from the BRST $s-$invariance of the S-matrix takes place outside the narrow limitation of gauge-invariant local operators and hence has less physical credibility. The result vindicates the Schwinger-Swieca charge-screening mechanism behind the $H$-coupling of the Higgs model. Since Schwinger's screening conjecture referred to massive QED, the coupling of a Hermitian field which first appeared in the somewhat metaphoric veil of the Higgs model was by all means an enrichment of possibilities of couplings of massive vectormesons with matter. Apart from the omnipresent stringlocal scalar Hermitian escorts \phi ,~$all couplings of massive vectormesons to matter (including the $H -coupling) can be removed by applying Occam's razor and therefore are hardly of the foundational significance which the mass-creating Higgs mechanism attributes to them. The appearance of a massive gluonium state with the quantum numbers attributed to the Higgs particle in a system of self-interacting massive vectormesons would change the conceptual situation. In such cases of such bound state problems (e.g. hadrons in terms of quarks) one usually looks for phenomenological description and the $H$-coupling may well be such a description. Significant progress from the new ideas is expected in the hitherto unsolved problems which hide behind the perturbative logarithmic $m\rightarrow 0$ infrared divergencies of stringlocal fields. Here the Hilbert space positivity is expected to show its full strength; the induction mechanism for the lowest order self-interactions between massive vectormeson relates the possible as independent presumed couplings; for equal masses the trilinear first order couplings must be antisymmetric and the second order induction is expected to have a Lie-algebra structure. In other words the properties which hitherto entered via the classical geometry of fibre-bundles and quantization into the QFT of $s=1$ interactions (and finally took the form of operator BRST gauge theory) can obtained without classical crutches alone from the Hilbert space positivity of the SLF formulation of interactions between massive vectormesons. There is another important attention-demanding aspect of the Hilbert space based perturbative stringlocal formulation of $s\geq 1$ interactions. According to the LSZ reduction formula the scattering amplitudes are simply the on-shell restrictions of the momentum space correlation functions. Hence unless a special mechanism intervenes, one expects that the off-shell high energy degree is passed to the scattering amplitudes. The "peeling property", i.e. the disposal of the leading short distance singularities of the form of divergencies $\partial V$ in the adiabatic limit is easily overlooked in momentum space, in particular since they cannot be encoded into Feynman graphs and already occur for tree-contributions to the scattering amplitude. The validity of graphical representations (Feynman graphs) is limited to pointlike renormalizable interactions for $s<1;~ formally they also hold for the gauge setting in Krein space except that scattering amplitudes are already outside the range of gauge invariant local observables. This raises the important question whether the more popular functional formalism (instead of an operator formulation) remains reliable for $s\geq 1$ interactions. Looking back at the history of renormalized perturbation theory it is conspicuous that there has been no noteworthy conceptual investment with computational consequences since the post wwII discovery of renormalized perturbation theory in the wake of Lagrangian quantization. Even the later extensions of these techniques to nonabelian gauge theories did not require significant new conceptual investments; as before, prescriptions for the removal of infinities and consistency checks for the so-obtained finite expressions was all what was needed. Later conceptual progress which allowed to derive these results from the iterative perturbative implementation of the foundational causal localization principle was hardly noticed by the majority of practicing particle theorists. The "think as (or often after) you pull up your sleeves and compute along" attitude became the motto of "robust" and very successful particle theory to the extend that it led to extreme claims e.g. that any physically inspired mathematically correct calculation will find its material realization in one of the imagined parallel universes. It seems that this successful period in particle theory had reached its apogee at the time of the discovery of the Higgs mechanism. Identical Lagrangian manipulations which led to identical conclusions were obtained in at least four independent publications at approximately the same time (a rather unique event in the history of particle theory), and while on the experimental side there has been a steady progress mostly confirming theoretical ideas, particle theory went through a already 40 year lasting period of impressive conceptual stagnation. The concepts and even their formulation (including the mass-giving "fattening" of photons) have remained identical, and completely reasonable ideas as Schwinger's charge screening for massive vectormesons and its mathematical derivation from first principles were allowed to vanish in the maelstrom of time. This is the clearest indication that the conceptual reservoir of the post wwII particle theory has been used up and new investments are urgently needed. The present proposal of extending the renormalizable Hilbert space operator setting of perturbation theory to higher spins $s\geq 1,~$and confront the localization problems caused by maintaining the Hilbert space positivity head-on, is meant as a contribution to this goal. Finally some remarks about the history behind the present ideas may be helpful. Although it may be seen as a new startup of old ideas of DeWitt \cite{DeWitt} and Mandelstam \cite{Man}, this is not the way in which it arose. The booster was rather our solution of the old problem of the field theoretic content of Wigner's infinite spin positive energy representation of the Poincar\'{e} group \cite{MSY}. Although it was already clear by 1970 that this representation does not admit an associated pointlike field description \cite{Y}, the first clue in what direction to look came from the application of \textit{modular localization} to the positive energy Wigner representation \cite{MSY}; the remaining problem of converting this observation into quantum fields associated to Wigner's infinite spin representation was solved in the cited work. Whereas in that case the string localization was endemic (every field, including composites has a noncompact localization), its use for finite helicity representation was only necessary if one were to use potentials instead of field strengths. The short distance scale dimension reducing role of string-localization was noted in \cite{MSY} and the first remarks about how to use this in order to convert nonrenormalizable pointlike couplings into power-counting renormalizable interactions of stringlocal fields can be found in \cite{Rio} and \cite{Bros . Meanwhile there exist strong indications that the noncompact localizable infinite spin Wigner "stuff" is more than a booster for directing the conceptual attention to string-localization. Its solely noncompact modular localization (no pointlike composites) leads to "inert matter"; accepting the standard picture about counter registration of particles, a particle event is interpreted as a reduction of a field state into a quasi-local counter-centered particle wave function\ \cite{Haag}. In case of the inherently noncompact Wigner stuff the localization of fields and particles are both noncompact and the situation of a counter registration can not be realized. However the stability and the coupling to gravitation which are properties shared by all positive energy representations; for this reason the Wigner stuff is the ideal candidate of "darkness". In fact interacting confined matter and free noncompact Wigner stuff stand in a conceptually interesting contrast \cite{dark}. Although they share the \textit irreducibility} of their zero mass strings (which accounts for the darkness of free strings and the confinement of interacting strings) whereas the strings of electrically charged fields of QED remain reducible \cit {stringlocal}. \begin{acknowledgement} I am indebted to Jens Mund for numerous discussions and I am looking forward to the publication of his mathematical-conceptual support for the new ideas as well as projected joint publications. I also thank Karl-Henning Rehren for a critical reading of the manuscript. Special thanks go to Raymond Stora who accompanied this project with interest and encouragement and last not least for sharing his over more than during 5 decades acquired critical wisdom about gauge theory. \end{acknowledgement}
\section{Introduction} According to the $AdS/CFT$ correspondence this superconformal field theory is dual to the eleven dimensional supergravity on $AdS_4 \times S_7$. Apart from a constant closed 7-form on $S^7$, $AdS_4 \times S_7 \sim SO(2,3)\times SO(1,2)/ SO(8)\times SO(7) \subset OSp(8|4)/SO(1,3) \times SO(7)$. So, the dual superconformal field theory to the eleven dimensional supergravity on $AdS_4 \times S_7$ has $OSp(8|4)$ realized as $\mathcal{N} = 8$ supersymmetry. This theory also has eight gauge valued scalar fields, sixteen physical fermions and the gauge fields of this theory do not have any on-shell degrees of freedom. All these properties are satisfied by a theory called the BLG theory \cite{1, 2, 3, 4, 5}. The BLG theory is based on gauge symmetry generated by a Lie 3-algebra rather than a Lie algebra. So, far the only know example of a Lie 3-algebra is $SO(4) \sim SU(2) \times SU(2)$, and it corresponds to two M2-branes. It has not been possible to increase the rank of the gauge group. It has been possible to construct a superconformal gauge theory called the ABJM theory \cite{abjm, ab, ab1, ab2}. The ABJM theory only has $\mathcal{N} = 6$ supersymmetry. However, it considers with the $BLG$ theory for the only known example of the Lie 3-algebra and so its supersymmetry is expected to get enhanced to full $\mathcal{N} = 8$ supersymmetry \cite{abjm2}. The gauge sector is described by two Chern-Simons theories with levels $k$ and $-k$. The matter fields in the ABJM theory are in the bi-fundamental representation of the gauge group $U(N)_k \times U(N)_{-k}$ and the gauge fields are in the adjoint representation. The ABJM theory has been studied in $\mathcal{N} =1$ and $\mathcal{N} =2$ superspace formalism \cite{6, 7, 8}. The ABJM theory has also been studied in harmonic superspace \cite{ahs, mir4}. However, in this paper, we will analyse the ABJM theory in $\mathcal{N} =1$ superspace formalism. The BRST and the anti-BRST symmetries for the ABJM theory have been studdied in both linear and non-linear gauges \cite{abm}. The infinitesimal BRST transformations have been generalized to finite field dependent BRST (FFBRST) originally in \cite{jm} and further generalized to construct finite field dependent anti-BRST (FFanti-BRST) transformations in \cite{antifbrst} Similar generalizations have also been made recently in \cite{lav,lav1}. This is done by first making the infinitesimal global parameter occurring in the BRST or the anti-BRST transformations depend on fields occurring in the theory. Then this field dependent parameter is integrated to obtain the FFBRST and anti-FFBRST transformations. Even though, these finite transformations are a symmetry of the quantum action, they are not a symmetry of the functional measure. They can thus be used to relate a theory in one gauge to the same theory in a different gauge \cite{lav}-\cite{ff1}. So, FFBRST transformations can be used to overcome a problem that a theory suffers from in a particular gauge. This can be done by first calculating the required quantity in a gauge in which that problem does not exist, and then using the FFBRST transformation to transform it to the required gauge. Thus, in Yang-Mills theory, FFBRST transformations have been used for obtaining the propagator in Coulomb gauge from the generating function in the Lorentz gauge \cite{ffb1}. The gauge-fixing and ghost terms corresponding to Landau and maximal Abelian gauge for the Cho-Faddeev-Niemi decomposed $SU(2)$ theory have also been generated using FFBRST transformation \cite{sud}. However, the linear and non-linear gauges of perturbative quantum gravity are connected at both classical and quantum level through FFBRST formulation \cite{sud1}. The quantum gauge freedom described by gaugeon formalism has also been studied for quantum gravity \cite{sud2} as well as for Higgs model \cite{sud3} utilizing FFBRST technique. The FFBRST transformations are also studied in the context of lattice gauge theory \cite{rbs1} and relativistic point particle model \cite{rbs}. The FFBRST transformation is used to relate the Gribov-Zwanziger theory to Yang-Mills theory in Landau gauge \cite{gz}. The problem of formulating the Gribov-Zwanziger theory beyond the Landau gauge is very delicate matter and substantial progress has been made recently towards the study of this problem \cite{lav2,lav3}. Thus, FFBRST transformations may give us an idea about the non-perturbative effects in a theory. This is very important from the M-theory point of view. This is because we may be able to understand the physics of multiple M5-branes by analysing non-perturbation effects in the ABJM theory \cite{m5}-\cite{m512}. The FFBRST transformations for the BLG theory has already been studied \cite{fs}. However, this limits the analysis to two M2-branes. If we want to analyse similar effects for multiple M2-branes, we need to analyse a similar system for ABJM theory. It may be noted that in analysing the FFBRST symmetry for the ABJM theory, we will need to introduce two finite field dependent paremeters, which correspond to the gauge symmetries generated by $U(N)_k \times U(N)_{-k}$. As the matter fields transform in bi-fundamental representation of this gauge group, the matter sector mixes these two finite field dependent paremeters. Thus, we need to generalize the ordinary FFBRST symmetry, to apply it on the ABJM theory. This is what we aim to do in this paper. The paper is organized as follows. In Sec. 2, we discuss the preliminaries about ABJM theory in ${\cal N}=1$ superspace. The BRST symmetry for various gauges are presented in Sec. 3. The FFBRST transformation for ABJM theory is developed in Sec. 4. In Sec. 5, we relate two arbitrary gauges of ABJM theory using FFBRST transformation. \section{ABJM Theory in ${\cal N}=1$ Superspace} In this section we analyse ABJM theory on ${\cal N}=1$ superspace. For this purpose, we begin with the Chern-Simons Lagrangian densities $\mathcal{L}_{CS}$, $\tilde{\mathcal{L}}_{CS}$ with gauge group's $U(N)_k$ and $U(N)_{-k}$ on ${\cal N}=1$ superspace defined by \begin{eqnarray} \mathcal{L}_{CS}& =& \frac{k}{2\pi} \int d^2 \, \theta \, \, \mbox{Tr} \left[ \Gamma^a \omega_a + \frac{i}{3} [\Gamma^a, \Gamma^b]_{ } D_b \Gamma_a + \frac{1}{3} [\Gamma^a,\Gamma^b]_{ } [\Gamma_a, \Gamma_b]_{ } \right] , \nonumber \\ \tilde{\mathcal{L}}_{CS} &=& -\frac{k}{2\pi} \int d^2 \, \theta \, \, \mbox{Tr} \left[ \tilde{\Gamma}^a \tilde{\omega}_a + \frac{i}{3} [\tilde{\Gamma}^a, \tilde{\Gamma}^b]_{ } D_b \tilde{\Gamma}_a + \frac{1}{3} [\tilde{\Gamma}^a,\tilde{\Gamma}^b]_{ } [\tilde{\Gamma}_a, \tilde{\Gamma}_b]_{ }\right], \end{eqnarray} where $\omega_a$ and $\tilde \omega_a$ have following expression: \begin{eqnarray} \omega_a &=& \frac{1}{2} D^b D_a \Gamma_b - i [\Gamma^b, D_b \Gamma_a]_{ } - \frac{2}{3} [ \Gamma^b , [ \Gamma_b, \Gamma_a]_{ }]_{ },\nonumber \\ \tilde\omega_a &=& \frac{1}{2} D^b D_a \tilde\Gamma_b -i [\tilde\Gamma^b, D_b \tilde\Gamma_a]_{ } - \frac{2}{3} [ \tilde\Gamma^b, [ \tilde\Gamma_b, \tilde\Gamma_a]_{ } ]_{ }, \end{eqnarray} with the super-derivative $D_a$ defined by $ D_a = \partial_a + (\gamma^\mu \partial_\mu)^b_a \theta_b. $ In the component form the super-gauge connections $\Gamma_a$ and $\tilde \Gamma_a$ are described by \begin{eqnarray} \Gamma_a = \chi_a + B \theta_a + \frac{1}{2}(\gamma^\mu)_a A_\mu + i\theta^2 \left[\lambda_a - \frac{1}{2}(\gamma^\mu \partial_\mu \chi)_a\right], \nonumber \\ \tilde\Gamma_a = \tilde\chi_a + \tilde B \theta_a + \frac{1}{2}(\gamma^\mu)_a \tilde A_\mu + i\theta^2 \left[\tilde \lambda_a - \frac{1}{2}(\gamma^\mu \partial_\mu \tilde\chi)_a\right]. \end{eqnarray} The Lagrangian density of the matter fields is given by \begin{eqnarray} \mathcal{L}_{M} =\frac{1}{4} \int d^2 \, \theta \, \, \mbox{Tr} \left[ [\nabla^a_{(X)} X^{I \dagger} \nabla_{a (X)} X_I ] + [\nabla^a_{(Y)} Y^{I \dagger} \nabla_{a (Y)} Y_I ] + \frac{16\pi}{k} \mathcal{V}_{ } \right] , \end{eqnarray} where \begin{eqnarray} \nabla_{(X)a} X^{I } &=& D_a X^{I } + i \Gamma_a X^I - i X^I \tilde\Gamma_a , \nonumber \\ \nabla_{(X)a} X^{I \dagger} &=& D_a X^{I \dagger} + i \tilde\Gamma_a X^{I \dagger}- i X^{I \dagger} \Gamma_a, \nonumber \\ \nabla_{(Y)a} Y^{I } &=& D_a Y^{I } + i \tilde\Gamma_a Y^I- i Y^I \Gamma_a , \nonumber \\ \nabla_{(Y)a} Y^{I \dagger} &=& D_a Y^{I \dagger} + i \Gamma_a Y^{I \dagger} - i Y^{I \dagger} \tilde\Gamma_a. \end{eqnarray} Now, the gauge invariant Lagrangian density for ABJM theory with the gauge group $U(N)_k \times U(N)_{-k} $ on ${\cal N}=1$ superspace is given by, \begin{equation} { \mathcal{L}_c} = \mathcal{L}_{M} + \mathcal{L}_{CS} - \tilde{\mathcal{L}}_{CS}. \end{equation} The gauge transformations are given by \begin{eqnarray} \delta \,\Gamma_{a} = \nabla_a \xi, && \delta \, \tilde\Gamma_{a} =\tilde\nabla_a \tilde \xi, \nonumber \\ \delta \, X^{I } = i \xi X^{I } - iX^{I } \tilde \xi, && \delta \, X^{I \dagger } = i \tilde \xi X^{I \dagger } - i X^{I \dagger } \xi, \nonumber \\ \delta \, Y^{I } = i \tilde \xi Y^{I } -iY^{I } \xi, && \delta \, Y^{I \dagger } = i \xi Y^{I \dagger } - i Y^{I \dagger } \tilde\xi, \end{eqnarray} with the local parameters $\xi$ and $\tilde\xi$. Here, the super-covariant derivatives $\nabla_a$ and $\tilde\nabla_a$ are defined by \begin{equation} \nabla_a =D_a-i\Gamma_a, \ \ \ \tilde\nabla_a =D_a-i\tilde\Gamma_a. \end{equation} Not all the degrees of freedom of this theory are physical as it is invariant under gauge transformations. \section{BRST Symmetry} In this section we will review the BRST symmetry for the ABJM theory in the ${\cal N}=1$ superspace. Being gauge invariant, ABJM theory cannot be quantized without getting rid of these unphysical degrees of freedom. This is done by fixing the following gauge, \begin{eqnarray} G_1 \equiv D^a \Gamma_a =0,\ \ \tilde G_1 \equiv D^a \tilde{\Gamma}_a =0. \end{eqnarray} These gauge fixing conditions are incorporated at a quantum level by adding a gauge fixing term $\mathcal{L}_{gf }$ and a ghost term $\mathcal{L}_{gh }$ to the original classical Lagrangian. Here the gauge fixing term is given by \begin{equation} \mathcal{L}_{gf} = \int d^2 \, \theta \, \, \mbox{Tr} \left[ib (D^a \Gamma_a) + \frac{\alpha}{2}b b - i \tilde{b} (D^a \tilde{\Gamma}_a) - \frac{\alpha}{2}\tilde{b} \tilde{ b} \right] , \end{equation} where $b $ and $\tilde b $ are the Nakanishi-Lautrup auxiliary fields. The Faddeev-Popov ghost term is given by \begin{equation} \mathcal{L}_{gh} = \int d^2 \, \theta \, \, \mbox{Tr} \left[ i\bar{c} D^a \nabla_a c -i \tilde{\bar{c}} D^a \tilde{\nabla}_a \tilde{c} \right] . \end{equation} The sum of the original Lagrangian density with the gauge fixing and ghost terms is invariant under the following BRST transformations \begin{eqnarray} \delta_b \,\Gamma_{a} = \nabla_a c \ \Lambda, && \delta_b\, \tilde\Gamma_{a} =\tilde\nabla_a \tilde c \ \tilde \Lambda, \nonumber \\ \delta_b\,c = - {[c ,c ]}_ { } \Lambda, && \delta_b \,\tilde{ {c}} = - [\tilde{ {c}} , \tilde c ]_{ } \tilde \Lambda, \nonumber \\ \delta_b \,\bar{c} = b \ \Lambda, && \delta_b \,\tilde {\bar c} = \tilde b \ \tilde \Lambda, \nonumber \\ \delta_b \,b =0, &&\delta_b \, \tilde b = 0, \nonumber \\ \delta_b \, X^{I } = i \Lambda c X^{I } - iX^{I } \tilde c \tilde \Lambda, && \delta_b \, X^{I \dagger } = i \tilde \Lambda \tilde c X^{I \dagger } - i X^{I \dagger } c \ \Lambda, \nonumber \\ \delta_b \, Y^{I } = i \tilde \Lambda \tilde c Y^{I } -iY^{I } c \ \Lambda, && \delta_b \, Y^{I \dagger } = i \Lambda c Y^{I \dagger }\ - i Y^{I \dagger } \tilde c \ \tilde \Lambda, \label{brstl} \end{eqnarray} where $\Lambda$ and $ \tilde \Lambda$ are the infinitesimal anticommuting parameters of transformation. Now we analyse ABJM theory in non-linear gauge and therefore we define the Lagrangian density as follows \begin{eqnarray} {\cal L}_{NL} &=& {\cal L}_c+ \int d^2\theta \ \mbox{Tr} \bigg[ \frac{\alpha}{2} b ^2 + i b D^a \Gamma_a -iD^a \bar{c} \nabla_ac - \frac{i}{2} D^a \Gamma_a [\bar{c} ,c ] \nonumber \\ &+& \frac{\alpha}{8} [\bar{c} ,c ]^2 - \frac{\alpha}{2}b [\bar{c} ,c ] + iD^a \tilde{\bar{c}} \nabla_a\tilde{c} - \frac{\alpha}{2} \tilde{b} ^2 - i\tilde{b} D^a \tilde{\Gamma}_a \nonumber \\ &+& \frac{i}{2} D^a \tilde{\Gamma}_a [\tilde{\bar{c}} ,\tilde{c} ]- \frac{\alpha}{8} [\tilde{\bar{c}} ,\tilde{c} ]^2 +\frac{\alpha}{2} \tilde{b} [\tilde{\bar{c}} , \tilde{c} ] \bigg] .\label{lag} \end{eqnarray} We notice that the above Lagrangian density can be obtained by shifting the Nakanishi-Lautrup auxiliary fields as follows \begin{equation} b \rightarrow b -\frac{1}{2} [\bar c , c ], \ \ \tilde b \rightarrow \tilde b -\frac{1}{2} [\tilde{\bar c }, \tilde c ]. \end{equation} The BRST transformation, under which the effective action in non-linear gauge (\ref{lag}) is invariant, is given by \begin{eqnarray} \delta_b \,\Gamma_{a} = \nabla_a c \ \Lambda, && \delta_b\, \tilde\Gamma_{a} =\tilde\nabla_a \tilde c \ \tilde\Lambda, \nonumber \\ \delta_b \,c = -\frac{1}{2} {[c ,c ]}_ { }\ \Lambda, && \delta_b \,\tilde{ {c}} = - \frac{1}{2} [\tilde{ {c}} , \tilde c ]_{ } \ \tilde\Lambda, \nonumber \\ \delta_b \,\bar{c} = b \ \Lambda -\frac{1}{2}[\bar c , c ] \Lambda, && \delta_b \,\tilde {\bar c} = \tilde b \ \tilde\Lambda -\frac{1}{2}[\tilde {\bar c} , \tilde c ]\ \tilde\Lambda, \nonumber \\ \delta_b \,b =-\frac{1}{2}[c , b ]\Lambda- \frac{1}{8}[ [c ,c ] , \bar c ]\Lambda, &&\delta_b \, \tilde b = -\frac{1}{2}[\tilde c , \tilde b ] \tilde\Lambda- \frac{1}{8}[ [\tilde c , \tilde c ], \tilde{\bar c }]\tilde\Lambda, \nonumber \\ \delta_b \, X^{I } = i \Lambda c X^{I }- iX^{I } \tilde c \ \tilde\Lambda, && \delta_b \, X^{I \dagger } = i \tilde\Lambda \tilde c X^{I \dagger }\ - i X^{I \dagger } c \ \Lambda, \nonumber \\ \delta_b \, Y^{I } = i\tilde\Lambda \tilde c Y^{I }\ -iY^{I } c \ \Lambda, && \delta_b \, Y^{I \dagger } = i\Lambda c Y^{I \dagger } \ - i Y^{I \dagger } \tilde c \ \tilde\Lambda. \label{brstnl} \end{eqnarray} Remarkably, the effective action is also found invariant under the another set of BRST symmetry (called as anti-BRST transformation) where roles of ghost and anti-ghost fields are interchanged. The anti-BRST transformation is written by \begin{eqnarray} \delta_{ab} \,\Gamma_{a} = \nabla_a \bar c \ \bar\Lambda, && \delta_{ab}\, \tilde\Gamma_{a} =\tilde\nabla_a \tilde{\bar c }\ \tilde{\bar \Lambda}, \nonumber \\ \delta_{ab} \,\bar c = - \frac{1}{2}{[\bar c , \bar c ]}_ { }\ \bar\Lambda, &&\delta_{ab} \,\tilde{\bar {c}} = - \frac{1}{2} [\tilde{\bar {c}} , \tilde {\bar c} ]_{ }\ \tilde{\bar \Lambda}, \nonumber \\ \delta_{ab} \, {c} = -b \ \bar\Lambda-\frac{1}{2}[\bar c , c ]\ \bar\Lambda, &&\delta_{ab} \,\tilde { c} = -\tilde b \ \tilde{\bar \Lambda} -\frac{1}{2}[\tilde {\bar c} , \tilde c ]\ \tilde{\bar \Lambda}, \nonumber \\ \delta_{ab} \,b =-\frac{1}{2}[ \bar c , b ]\ \bar\Lambda +\frac{1}{8}[ [\bar c , \bar c ], c ]\ \bar\Lambda, &&\delta_{ab} \, \tilde b = -\frac{1}{2} [ \tilde {\bar c} , \tilde b ]\ \tilde{\bar \Lambda}+\frac{1}{8} [[\tilde {\bar c} , \tilde {\bar c} ], \tilde{ c} ]\ \tilde{\bar \Lambda}, \nonumber \\ \delta_{ab} \, X^{I } = i \bar\Lambda \bar c X^{I }\ - iX^{I } \tilde {\bar c} \ \tilde{\bar \Lambda}, && \delta_{ab} \, X^{I \dagger } = i \tilde{\bar \Lambda} \tilde {\bar c} X^{I \dagger }\ - i X^{I \dagger } \bar c \ \bar\Lambda, \nonumber \\ \delta_{ab} \, Y^{I } = i \tilde{\bar \Lambda} \tilde{\bar c} Y^{I } \ -iY^{I } \bar c \ \bar\Lambda, && \delta_{ab} \, Y^{I \dagger } = i \bar\Lambda \bar c Y^{I \dagger }\- i Y^{I \dagger } \tilde{\bar c} \ \tilde{\bar \Lambda}. \end{eqnarray} The above BRST and anti-BRST transformations satisfy the following algebra: \begin{eqnarray} \delta_{b}^2=0,\ \ \delta_{ab}^2 =0,\ \ \delta_{b}\delta_{ab} +\delta_{ab} \delta_{b}=0. \end{eqnarray} With these BRST and anti-BRST transformations the Lagrangian density (\ref{lag}) can also be expressed as \begin{eqnarray} {\cal L}_{NL} &=&{\cal L}_c+ \frac{i}{2}\delta_{b}\delta_{ab}\int d^2\theta\ \mbox{Tr} \left[\Gamma_a \Gamma^a -\tilde \Gamma_a\tilde\Gamma^a -i\alpha \bar c c +i\alpha \tilde{\bar c }\tilde c \right] ,\nonumber\\ &=&{\cal L}_c-\frac{i}{2} \delta_{ab} \delta_{b}\int d^2\theta\ \mbox{Tr}\left[\Gamma_a \Gamma^a -\tilde \Gamma_a\tilde\Gamma^a -i\alpha \bar c c +i\alpha \tilde{\bar c }\tilde c \right] . \end{eqnarray} \section{Finite Field Dependent Transformation} In this section we construct finite field dependent BRST transformation \cite {jm} of ABJM theory in ${\cal N}=1$ superspace. To do that we first define two sets of generaric fields as $\Phi^{i }_L (x, \kappa) \equiv (\Gamma_{a}, X^I, Y^I, c, \overline{c}, b )$ and $\Phi^{i }_R (x, \kappa) \equiv (\tilde\Gamma_{a}, \tilde X^I, \tilde Y^I, \tilde c, \tilde{\overline{c}}, b)$ , here the parameter $\kappa: 0\le \kappa \le 1$. Here ${\Phi^i}_L (x, 0 ), {\Phi^i}_R (x, 0 )$ are the initial fields and $ {\Phi^i}_L (x, 1), {\Phi^i}_R (x, 1 )$ are the transformed fields. The infinitesimal but field dependent BRST transformations can be written as \cite{jm} \begin{eqnarray} \frac{ d}{d \kappa}{\Phi^i}_L (x, \kappa ) &=& s {\Phi^i}_L (x )\ \epsilon_L [{\Phi}_L (x,\kappa )], \nonumber \\ \frac{ d}{d \kappa}{\Phi^i}_R (x, \kappa ) &=& s {\Phi^i}_R (x )\ \epsilon_L [{\Phi}_R (x,\kappa )]. \end{eqnarray} where $ \epsilon_L [{\Phi}_L (x)]$ and $\epsilon_R [{\Phi}_R (x)]$ are infinitesimal field dependent parameters. Now integrating the above equation from $ \kappa=0$ to $\kappa=1$, we get the FFBRST transformation, \begin{eqnarray} {\Phi^i}_L (x, 1) = {\Phi^i}_L (x, 0) + s {\Phi^i}_L (x) \Theta_L [{\Phi}_L (x)], \nonumber \\ {\Phi^i}_R (x, 1) = {\Phi^i}_R (x, 0) + s {\Phi^i}_R (x) \Theta_R [{\Phi}_R (x)], \end{eqnarray} where finite field dependent parameters are \begin{eqnarray} \Theta_L [{\Phi}_L (x)] &= &\int_0^1 d\kappa\ \epsilon_L [{\Phi}_L (x,\kappa )],\nonumber\\ \Theta_R [{\Phi}_R (x)] &= &\int_0^1 d\kappa\ \epsilon_R [{\Phi}_R (x,\kappa )].\label{para} \end{eqnarray} Furthermore these finite parameters are calculated \cite{jm} as, \begin{eqnarray} \Theta_L [{\Phi}_L (x)] &=& \epsilon_L [{\Phi}_L (x)] \frac{ \exp F_L [{\Phi}_L (x)] -1}{F_L [{\Phi}_L (x)]}, \nonumber \\ \Theta_R [{\Phi}_R (x)] &=& \epsilon_R [{\Phi}_R (x)] \frac{ \exp F_R [{\Phi}_R (x)] -1}{F_R [{\Phi}_R (x)]}, \end{eqnarray} where \begin{eqnarray} F_L &=& \sum_i\frac{ \delta \epsilon_L [\Phi_L (x)]}{\delta \Phi^i_L (x)} s \Phi^i_L (x), \nonumber \\ F_R &=& \sum_i\frac{ \delta \epsilon_R [\Phi_R (x)]}{\delta \Phi^i_R (x)} s \Phi^i_R (x). \label{ref} \end{eqnarray} Now the FFBRST transformations in the linear gauge are given by \begin{eqnarray} \delta_b \,\Gamma_{a} = \nabla_a c \ \Theta_L, && \delta_b\, \tilde\Gamma_{a} =\tilde\nabla_a \tilde c \ \Theta_R, \nonumber \\ \delta_b\,c = - {[c ,c ]}_ { } \Theta_L, && \delta_b \,\tilde{ {c}} = - [\tilde{ {c}} , \tilde c ]_{ } \Theta_R, \nonumber \\ \delta_b \,\bar{c} = b \ \Theta_L, && \delta_b \,\tilde {\bar c} = \tilde b \ \Theta_R, \nonumber \\ \delta_b \,b =0, &&\delta_b \, \tilde b = 0, \nonumber \\ \delta_b \, X^{I } = i \Theta_Lc X^{I } - iX^{I } \tilde c \Theta_R, && \delta_b \, X^{I \dagger } = i \Theta_R \tilde c X^{I \dagger } - i X^{I \dagger } c \ \Theta_L, \nonumber \\ \delta_b \, Y^{I } = i \Theta_R \tilde c Y^{I } -iY^{I } c \ \Theta_L, && \delta_b \, Y^{I \dagger } = i \Theta_L c Y^{I \dagger }\ - i Y^{I \dagger } \tilde c \ \Theta_R. \end{eqnarray} and the FFBRST transformations in the non-linear gauge are given by \begin{eqnarray} \delta_b \,\Gamma_{a} = \nabla_a c \ \Theta_L, && \delta_b\, \tilde\Gamma_{a} =\tilde\nabla_a \tilde c \ \Theta_R, \nonumber \\ \delta_b \,c = -\frac{1}{2} {[c ,c ]}_ { }\ \Theta_L, && \delta_b \,\tilde{ {c}} = - \frac{1}{2} [\tilde{ {c}} , \tilde c ]_{ } \ \Theta_R, \nonumber \\ \delta_b \,\bar{c} = b \ \Theta_L -\frac{1}{2}[\bar c , c ]\Theta_L, && \delta_b \,\tilde {\bar c} = \tilde b \ \Theta_R-\frac{1}{2}[\tilde {\bar c} , \tilde c ]\ \Theta_R, \nonumber \\ \delta_b \,b =-\frac{1}{2}[c , b ]\Theta_L - \frac{1}{8}[ [c ,c ] , \bar c ]\Theta_L, &&\delta_b \, \tilde b = -\frac{1}{2}[\tilde c , \tilde b ] \Theta_R - \frac{1}{8}[ [\tilde c , \tilde c ], \tilde{\bar c }]\Theta_R, \nonumber \\ \delta_b \, X^{I } = i \Theta_L c X^{I }- iX^{I } \tilde c \ \Theta_R, && \delta_b \, X^{I \dagger } = i \Theta_R \tilde c X^{I \dagger }\ - i X^{I \dagger } c \ \Theta_L, \nonumber \\ \delta_b \, Y^{I } = i \Theta_R \tilde c Y^{I }\ -iY^{I } c \ \Theta_L, && \delta_b \, Y^{I \dagger } = i \Theta_L c Y^{I \dagger } \ - i Y^{I \dagger } \tilde c \ \Theta_R. \end{eqnarray} The Jacobians for path integral measures in the expression of generating functionals are given by \begin{eqnarray} {\cal D}{\Phi^i}_L =J_L[{\Phi}_L (\kappa)] {\cal D}{\Phi^i}_L (\kappa), && {\cal D}{\Phi^i}_R =J_R[{\Phi}_R (\kappa)] {\cal D}{\Phi^i}_R (\kappa). \end{eqnarray} So, the FFBRST transformations are not a symmetry of the generating functional. Now $J_L[{\Phi}_L (\kappa)]J_R[{\Phi}_R (\kappa)]$ can be replaced within the functional integral by $\exp\, ({iS_{1L}[{\Phi}_L (\kappa)]+iS_{1R}[{\Phi}_R (\kappa)]})$, if the following equations are satisfied, \begin{eqnarray} \int d^2\theta \ \mbox{Tr} \left[\frac{1}{J_L (\kappa )}\frac{d J_L (\kappa )}{d\kappa} -i\frac{dS_{1L} }{d\kappa}\right] =0, \nonumber \\ \int d^2\theta \ \mbox{Tr} \left[\frac{1}{J_R(\kappa )}\frac{d J_R (\kappa )}{d\kappa} -i\frac{dS_{1R}}{d\kappa}\right] =0. \label{mcond} \end{eqnarray} The infinitesimal changes in Jacobian's are given by \begin{eqnarray} \frac{1}{J_L (\kappa )}\frac{d J_L (\kappa )}{d\kappa} &=& -\int d^2\theta \ \mbox{Tr} \ \mathcal{A}_L , \nonumber \\ \frac{1}{J_R(\kappa)}\frac{dJ_R(\kappa)}{d\kappa}&=& -\int d^2\theta \ \mbox{Tr} \ \mathcal{A}_R, \label{jaceva} \end{eqnarray} where explicit expressions for $\mathcal{A}_L$ and $\mathcal{A}_R$ are given by \begin{eqnarray} \mathcal{A}_L &=& \left[ s \Gamma_a (x, \kappa) \frac{ \delta \epsilon_L [\Phi_L (x, k)]}{\delta \Gamma_a (x, k)} -s c (x, k) \frac{ \delta \epsilon_L [\Phi_L (x)]}{\delta c (x, k)} \right. \nonumber \\ &&\left. - s \overline c (x, k) \frac{ \delta \epsilon_L [\Phi_L (x, k)]}{\delta \overline c (x, k)} +s b (x, k) \frac{ \delta \epsilon_L [\Phi_L (x, k)]}{\delta b (x, k)} \right. \nonumber \\ &&\left. - s X^I (x, k) \frac{ \delta \epsilon_L [\Phi_L (x, k)]}{\delta X^I (x, k)} +s Y^I (x, k) \frac{ \delta \epsilon_L [\Phi_L (x, k)]}{\delta Y^I (x, k)} \right] , \nonumber \\ \mathcal{A}_R &=& \left[ s \tilde{\Gamma}_a (x,\kappa) \frac{ \delta \epsilon_R [\Phi_R (x, k)]}{\delta \tilde{\Gamma}_a (x, k)} -s \tilde c (x, k) \frac{ \delta \epsilon_R [\Phi_R (x)]}{\delta \tilde c (x, k)} \right. \nonumber \\ &&\left. - s \tilde{\overline c} (x, k) \frac{ \delta \epsilon_R [\Phi_R (x, k)]}{\delta \tilde{\overline c} (x, k)} +s \tilde b(x, k) \frac{ \delta \epsilon_R [\Phi_R (x, k)]}{\delta \tilde b (x, k)}\right. \nonumber \\ &&\left. - s X^{I\dag} (x, k) \frac{ \delta \epsilon_R [\Phi_R (x, k)]}{\delta X^{I\dag} (x, k)} +s Y^{I\dag} (x, k) \frac{ \delta \epsilon_R [\Phi_R(x, k)]}{\delta Y^{I\dag} (x, k)} \right]. \end{eqnarray} Here we note that the conditions (\ref{mcond}) provide us liberty to replace the Jacobians of path integral measure by the exponential of local functional within functional measure. Hence, the Jacobians amount a precise change in effective action of generating functional . One can also arrive at the same conclusion following the work in Ref. \cite{lav,lav1}. \section{ Relating Different Gauges} We will now use FFBRST to relate the generating functional in the linear gauge to the generating functional in the non-linear gauge. If the gauge fixing condition in the linear gauge is denoted by $G_{1L}[\Gamma_a], G_{1R}[ \tilde \Gamma_a]$ and the gauge fixing condition in the non-linear gauge is denoted by $G_{2L}[\Gamma_a], G_{1R}[ \tilde\Gamma_a]$, then, the linear BRST transformations of $G_{1L}[\Gamma_a], G_{1R}[ \tilde\Gamma_a]$ are denoted by $sG_{1L}, sG_{1R}$ and the the non-linear BRST transformations of $G_{2L}[\Gamma_a], G_{1R}[ \tilde\Gamma_a]$ are denoted by $sG_{2L}, s G_{2R}$. We define the infinitesimal field dependent parameter as follows \begin{eqnarray} \epsilon_L [\Phi_L] &=& i\gamma\int d^2\theta \ \mbox{Tr} \ \left[ \overline c \left( G_{ 1L} - G_{ 2L}\right)\right] , \nonumber \\ \epsilon_R [\Phi_R] &=& -i\gamma\int d^2\theta \ \mbox{Tr}\ \left[\tilde{\overline c} \left( G_{ 1R} - G_{ 2R}\right)\right] , \end{eqnarray} where $\gamma$ is an arbitrary constant parameter. Using definition given in (\ref{jaceva}), the change in Jacobian's can be calculated as follows, \begin{eqnarray} \frac{1}{J_L}\frac{dJ_L}{d\kappa} &=&i\gamma\int d^2\theta \ \mbox{Tr} \ \left[ b G_{ 1L} - b G_{ 2L} - ( sG_{ 1L} - sG_{ 2L})\bar c \right] ,\nonumber\\ &=&i\gamma\int d^2\theta \ \mbox{Tr} \ \left[ b G_{ 1L} - b G_{2L} +\bar c ( sG_{ 1L} - sG_{2L})\right] , \nonumber \\ \frac{1}{J_R}\frac{dJ_R}{d\kappa} &=&- i\gamma\int d^2\theta \ \mbox{Tr} \ \left[ \tilde b G_{ 1R} -\tilde b G_{ 2R} -( sG_{ 1R} - sG_{ 2R})\tilde{\bar c} \right] ,\nonumber\\ &=&- i\gamma\int d^2\theta \ \mbox{Tr} \ \left[ \tilde b G_{ 1R} - \tilde b G_{2R} +\tilde {\bar c} ( sG_{ 1R} - sG_{2R})\right] . \end{eqnarray} Furthermore, the local functionals $S_{1L}$ and $S_{1R}$ involved in the Jacobians are defined as \begin{eqnarray} S_{1L} &=&\int d^2\theta \ \mbox{Tr} \ [\xi_{1L} (\kappa) b G_{ 1L}+ \xi_{2L} (\kappa) b G_{ 2L} \nonumber \\ && +\xi_{3L}(\kappa) \overline c s G_{1L} +\xi_{4L}(\kappa)\overline c sG_{ 2L} ] , \nonumber \\ S_{1R} &=& \int d^2\theta \ \mbox{Tr} \ [\xi_{1R} (\kappa) \tilde b G_{ 1R}+ \xi_{2R} (\kappa) \tilde b G_{ 2R} \nonumber \\ && +\xi_{3R}(\kappa) \tilde{\overline c} s G_{1R} +\xi_{4R}(\kappa)\tilde{\overline c } sG_{ 2R} ] , \end{eqnarray} where $\xi_{iL}, \xi_{iR}, (i=1,2,3,4)$ are $\kappa$ dependent arbitrary parameters which satisfy the following initial boundary conditions, $ \xi_{iL} (\kappa =0) = \xi_{iR}(\kappa =0)=0 $. As all the fields depend on $\kappa$, so we can write \begin{eqnarray} \frac{dS_{1L}}{d\kappa}&=&\int d^2\theta \ \mbox{Tr} \ [\xi'_{1L} b G_{ 1L} +\xi_{1L} b s G_{ 1L}\epsilon_L \nonumber \\ &&+\xi_{2R} b s G_{2L} \epsilon_L +\xi'_{3L} \overline c s G_{1L} -\xi_{3L} b s G_{1L}\epsilon_L \nonumber\\ &&+\xi'_{4L} \overline c sG_{2_L} -\xi_{4L} b s G_{2L} \epsilon_L + \xi'_{2L} b G_{2L}] ,\nonumber\\ &=&\int d^2\theta \ \mbox{Tr} \ [\xi'_{1L} b G_{ 1L} + \xi'_{2L} b G_{ 2L}\nonumber \\ & &+\xi'_{3L} \overline c s G_{ 1L} +\xi'_{4L} \overline c sG_{ 2L} \nonumber\\ &&+ (\xi_{1L} -\xi_{3L} )b s G_{ 1L}\epsilon_L +(\xi_{2L} -\xi_{4L}) b s G_{2L} \epsilon_L] , \nonumber \\ \frac{dS_{1R}}{d\kappa}&=&\int d^2\theta \ \mbox{Tr} \ [\xi'_{1R} \tilde b G_{ 1R} +\xi_{1R} \tilde b s G_{ 1R}\epsilon_R \nonumber \\ &&+\xi_{2R} \tilde b s G_{2R} \epsilon_R +\xi'_{3R} \tilde{\overline c } s G_{1R} -\xi_{3R} \tilde b s G_{1R}\epsilon_R \nonumber\\ &&+\xi'_{4R} \tilde{\overline c} sG_{2_R} -\xi_{4R} \tilde b s G_{2R} \epsilon_R + \xi'_{2R} \tilde b G_{2R}] ,\nonumber\\ &=&\int d^2\theta \ \mbox{Tr} \ [\xi'_{1R} \tilde b G_{ 1R} + \xi'_{2R} \tilde b G_{ 2R}\nonumber \\ & &+\xi'_{3R} \tilde{\overline c} s G_{ 1R} +\xi'_{4R} \tilde{\overline c} sG_{ 2R} \nonumber\\ &&+ (\xi_{1R} -\xi_{3R} )\tilde b s G_{ 1R}\epsilon_R +(\xi_{2R} -\xi_{4R})\tilde b s G_{2R} \epsilon_R] . \end{eqnarray} The Jacobians of path integral measure can be written as $\exp\, ({iS_{1L}+ iS_{1R}})$, when the following equations are satisfied, \begin{eqnarray} &&\int d^2\theta \ \mbox{Tr} \ \left[ (\xi'_{1L}-\gamma) b G_{ 1L}+( \xi'_{2L} +\gamma ) b G_{ 2L} \right.\nonumber\\ &&\left. + ( \xi'_{3L} -\gamma ) \overline c sG_{1L} + (\xi'_{4L} +\gamma ) \overline c sG_{ 2L} \right.\nonumber\\ &&\left.+(\xi_{1L} -\xi_{3L} )b sG_{1L}\epsilon_L +(\xi_{2L} -\xi_{4L}) b sG_{2L}\epsilon_L \right] =0, \nonumber \\ &&\int d^2\theta \ \mbox{Tr} \ \left[ (\xi'_{1R}+\gamma) \tilde{b} G_{ 1R}+( \xi'_{2R} -\gamma ) \tilde bG_{ 2R} \right.\nonumber\\ &&\left. + ( \xi'_{3R} +\gamma ) \tilde{\overline c} sG_{1R} + (\xi'_{4R} -\gamma ) \tilde{\overline c} sG_{ 2R} \right.\nonumber\\ &&\left.+(\xi_{1R} -\xi_{3R} )\tilde b sG_{1R}\epsilon_R +(\xi_{2R} -\xi_{4R}) \tilde b sG_{2R}\epsilon_R \right] =0. \end{eqnarray} Equating the coeffiecients of the above expressions, and setting $\gamma =1$, we get $ \xi_{1L} = -\xi_{1R} = \kappa, \, \xi_{2L} = -\xi_{2R} = -\kappa, \, \xi_{3L} = -\xi _{3R} = \kappa, \, \xi_{4L} = -\xi_{4R} = -\kappa $. Now, if we add $ S_1 = S_{1L} (\kappa =1 ) + S_{1R} (\kappa =1 )$ to the original action in the non-linear gauge, we obtain the action in the linear gauge within a functional integral. So, under the FFBRST transformations the generating functional in the non-linear gauge transforms to the generating functional in the linear gauge. Similar computations can also been made following the work in Ref. \cite{lav,lav1} to show that the FFBRST transformation amounts finite change in gauge-fixing fermion of the path integral. \section{Conclusion} In this paper we analysed the FFBRST transformations for the ABJM theory in ${\cal N}=1$ superspace. We first have discussed the BRST for the ABJM theory in ${\cal N}=1$ superspace. Then we have integrated the infinitesimal parameter in the BRST transformations to obtain the FFBRST transformations. As the ABJM theory contains two Chern-Simons terms, we have constructed two finite parameters in the FFBRST transformations. These parameters are only mixed due to the matter terms. The BRST transformations of this theory have been studied in both linear as well as non-linear gauges. After analysing both the linear and non-linear BRST transformations, a finite field dependent version of these transformations has been developed. It has been shown that these two gauges can be related to each other via FFBRST transformations. Multiple D2-brane action has been derived from a multiple M2-brane action by means of a novel Higgs mechanism \cite{d2,sxsw, xswd, d21}. In this mechanism a vacuum expectation value is given to a scalar field which breaks the gauge group $U(N) \times U(N)$ down to its diagonal subgroup. The theory thus obtained is the Yang-Mills theory coupled to matter fields. It would be interesting to start with a gauge fixed ABJM theory in ${\cal N}=1$ superspace and use the novel Higgs mechanics to obtain the Yang-Mills theory coupled to matter fields. It would also be interesting to study the FFBRST transformations of the ABJM theory and the FFBRST transformations of the theory obtained after using the novel Higgs mechanics. It is expected that the FFBRST transformations for the ABJM theory will reduce to the the FFBRST transformations for the Yang-Mills theory coupled to matter fields. There is a dual symmetry to the BRST symmetry called the anti-BRST symmetry \cite{ an1, an2}. The finite field version of anti-BRST (anti-FFBRST) symmetry has also been studied \cite{antifbrst,alex}. It would be interesting to study this symmetry for the ABJM theory in ${\cal N}=1$ superspace. Furthermore, the ABJM theory in presence of a boundary has also been analysed \cite{abb}. In this theory new boundary degrees of freedom have to be added to make this theory gauge invariant. It would be interesting to analyse the FFBRST and anti-FFBRST symmetry for this ABJM theory in presence of a boundary. These transformations can be used to relate the generating functionals in case of ABJM theory in presence of a boundary.
\section{Introduction} This paper is concerned with random flights, that is with random motions in the Euclidean space $\mathbb{R}^d,d\geq 2$, performed at finite velocity (sometimes also called random evolutions). These continuous-time non-Markovian random motions have sample paths formed by straight lines, turning through any angle whatever, i.e. with uniformly distributed directions on the unit-radius sphere. These random flights are useful to describe concrete motions of particles in gases and for these reason have been investigated by many physicists and mathematicians over the years. The first ones who analyzed the random flights with a fixed number of deviations were K. Pearson and J.C. Kluyver and some years later, S. Chandrasekar wrote a long paper on this subject with applications to astronomy. Recent applications to the analysis of photon propagation in the Cosmic Microwave Background (CMB) radiation have been discussed in Reimberg and Abramo (2013). Furthermore, Martens {\it et al}. (2012) have shown that the probability law of planar random motions coincides with the explicit form of the van Hove function for the run-and-tumble model in two dimensions. This work gives an interesting and strong link between explicit solutions of the Lorentz model of electron conduction and the probability theory of random flights. For other possible applications of the random flights models, see Hughes (1995). The main object of our investigation is represented by the position $\{\underline{\bf X}_d(t),t>0\}$ reached after a fixed or a random number of deviations. The randomization of the number $\mathcal N(t)$ of steps and of the lengths of intermediate displacements has produced more flexible versions of random flights for which explicit distributions of $\{\underline{\bf X}_d(t),t>0\}$ have been obtained. Recently many papers have studied random motions with velocity $c>0,$ with uniformly distributed deviations at Poisson paced times in the plane and successively in the Euclidean space $\mathbb{R} ^d$ (see, for instance, Stadje, 1987, Masoliver {\it et al}., 1993, Kolesnik and Orsingher, 2005, for planar motions, Stadje, 1989, and Orsingher and De Gregorio, 2007, in $\mathbb{R}^d$). Franceschetti, (2007) established a relationship between the number of deviations $n$ and the dimension $d$ for which the conditional distributions $P\{\underline{\bf X}_d(t)\in \de \underline{\bf x}_d|\mathcal N(t)=n\}$ are uniform. The assumption that changes of direction are governed by a homogeneous Poisson process leads to explicit (conditional and unconditional) probability distributions of $\{\underline{\bf X}_d(t),t>0\}$ only for $d=2,4.$ The idea of assuming different forms for intermediate steps (displacements) has produced fruitful results in that the probability distribution of $\underline{\bf X}_d(t)$ can be explicitly produced for all spaces of dimension $d\geq 2.$ The random flight $(\underline\Theta,\underline\tau,\mathcal N(t)),$ is a triple where $\underline\Theta$ represents the ensemble of deviations during the time interval $[0,t]$, $\underline\tau$ is the vector of the lengths of the displacements for a fixed number $\mathcal N(t)$ number of changes of direction. These processes have been extensively investigated in the case where $\underline\tau$ has a Dirichlet distribution and $\underline\Theta$ is spherically uniform. The papers by Le Ca\"er (2010), (2011), De Gregorio and Orsingher (2012), De Gregorio (2014) and Letac and Piccioni (2014) are devoted to this case. Pogorui and Rodriguez-Dagnino (2011), (2013) have studied random flights where $\underline\tau$ has Erlang distribution, while Beghin and Orsingher (2010) obtained conditional distributions of $\{\underline{\bf X}_d(t),t>0\}$ where the steps $\underline\tau$ are Gamma random variables with parameter 2. The equations governing the unconditional probability distributions of the random flight $(\underline\Theta,\underline\tau,\mathcal N(t))$ have recently been obtained (see Garra and Orsingher, 2014). The case where $\underline\Theta$ has a specific non-uniform distribution has been studied by De Gregorio (2012). Motions on subspaces of $\mathbb{R}^d$ can be regarded as random flights with random velocities and are studied in De Gregorio and Orsingher (2012) and Pogorui and Rodriguez-Dagnino (2012). Asymptotic results for the position of these type of random walks have been obtained by Ghosh {\it et al}. (2014). We now describe in detail the structure of the random flights. Let us consider a particle or a walker which starts from the origin of $\mathbb{R}^d,d\geq 2$, and performs its motion with a constant velocity $c>0$. We indicate by $0=t_0<t_1<t_2<...<t_k<...$ the random instants at which the random walker changes direction and denote the length of time separating these instants by $\tau_k=t_k-t_{k-1}$, $k\geq 1$. Let $\mathcal N(t)=\sup\{k\geq 1:t_k\leq t\}$ be the (random) number of times in which the random motion changes direction during the interval $[0,t]$. If, at time $t>0$, $ \mathcal N(t)=n$, with $n\geq 1$, the random motion has performed $n+1$ displacements. We observe that ${\bf \underline\tau}_n=(\tau_1,...,\tau_n)\in S_n$, where $S_n$ represents the open simplex $$ S_n=\left\{(\tau_1,...,\tau_n)\in\mathbb R^n: 0<\tau_k<t-\sum_{j=0}^{k-1}\tau_j, k=1,2,...,n\right\},$$ with $\tau_0=0$ and $\tau_{n+1}=t-\sum_{j=1}^n\tau_j$. By $\underline\Theta_{n+1}=(\underline\theta_{d-1}^1,...,\underline\theta_{d-1}^k,...,\underline\theta_{d-1}^{n+1})$ we denote the vector of independent deviations performed at times $t_k,k=0,1,...,n.$ The $k$-th random variable $\underline\theta_{d-1}^k$ is a $(d-1)$-dimensional random variable with components $(\theta_{1}^k,...,\theta_{d-2}^k,\phi^k)$ distributed uniformly on the unit-radius sphere $\mathbb{S}_{1 }^{d-1}=\{\underline{\bf x}_d\in \mathbb R^d:||\underline{\bf x}_d||=1\}$. This means that the probability density function of $\underline{\theta}_{d-1}^k$ is equal to \begin{equation}\label{eq:jointdis1} \varphi(\underline{\theta}_{d-1}^k)=\frac{\Gamma(\frac d2)}{2\pi^{\frac d2}}\sin^{d-2}\theta_1^k\sin^{d-3}\theta_2^k\cdots \sin\theta_{d-2}^k, \end{equation} where $\theta_j^k\in[0,\pi],j\in\{1,...,d-2\},\,\phi^k\in[0,2\pi]$. Furthermore, $\tau_k$ and $\theta_j^k$ are independent for each $k$. Let us denote by $\{\underline{\bf X}_d(t),t>0\}$ the process representing the position reached, at time $t>0$, by the particle moving randomly according to the rules described above. The position $\underline{\bf X}_d(t)=(X_1(t),...,X_d(t)),$ is the main object of interest of the random flight and can be written, for $\mathcal N(t)=n,$ as \begin{equation}\label{eq:definition} \underline{\bf X}_d(t)=c\sum_{k=1}^{n+1}{\bf \underline{V}}_d^k\tau_k, \end{equation} where ${\bf\underline{V}}_{d}^k,k=1,2,...,n+1,$ are independent $d$-dimensional random vectors defined as follows $${\bf\underline{V}}_d^k=\left( \begin{array}{l} \sin\theta_1^k\sin\theta_2^k\cdot\cdot\cdot\sin\theta_{d-2}^k\sin\phi^{k} \\ \sin\theta_1^k\sin\theta_2^k\cdot\cdot\cdot\sin\theta_{d-2}^k\cos\phi^{k} \\ ...\\ \sin\theta_1^k\cos\theta_2^k\\ \cos\theta_1^k \end{array} \right)$$ and $(\theta_1^k,\theta_2^k,...,\theta_{d-2}^k,\phi^{k})$ are independent and identically distributed with density \eqref{eq:jointdis1}. In this work we analyze reflecting random walks defined by means of inversive geometry. In particular, this paper studies random flights reflecting on the $d$-dimensional sphere $\mathbb{S}_R^{d-1}=\{\underline{\bf x}_d:||\underline{\bf x}_d||=R\}$ with radius $R.$ The reflected processes are constructed by suitably manipulating the sample paths of the free random flight $\{\underline{\bf X}_d(t),t>0\}$. We assume that the part of the trajectories of the free random flight are reflected by inversion with respect to the sphere $\mathbb{S}_R^{d-1}$. This produces substantial changes in their form, because the segments of outside lying sample paths are converted into arcs of circle inside $\mathbb{S}_R^{d-1}$. The picture of the reflecting processes is therefore made up by straight lines (for sample paths which never crossed $\mathbb{S}_R^{d-1}$) and circular arcs (for the sample paths which performed excursions outside $\mathbb{S}_R^{d-1}$). The same approach is used by Aryasova {\it et al}. (2013) in the study of reflecting diffusion processes. The circular inversion permits us to write down the probability distribution $p_n^*(\underline{{\bf x}}_d,t)$ of the reflected process $\{\underline{{\bf X}}_d^*(t),t>0\}$ by exploiting the density function $p_n(\underline{{\bf x}}_d,t)$ of the free random flight $\{\underline{{\bf X}}_d(t),t>0\}.$ We show that \begin{align}\label{eq:condensintrod} p_{n}^*(\underline{{\bf x}}_d,t)&=\begin{cases} p_n(\underline{{\bf x}}_d,t){\bf 1}_{B_R^d}(\underline{{\bf x}}_d),& t\leq\frac{R}{c}\\ p_n(\underline{{\bf x}}_d,t){\bf 1}_{B_R^d}(\underline{{\bf x}}_d)+\frac{R^{2d}}{||\underline{{\bf x}}_d||^{2d}}p_n\left(R^2\frac{\underline{{\bf x}}_d}{||\underline{{\bf x}}_d||^2},t\right){\bf 1}_{C^d_{\frac{R^2}{ct},R}}(\underline{{\bf x}}_d),& t>\frac{R}{c}, \end{cases} \end{align} where $B_R^d$ is the ball in the space $\mathbb{R}^d$ with radius $R$, $C^d_{\frac{R^2}{ct},R}=\{\underline{{\bf x}}_d\in\mathbb R^d:\frac{R^2}{ct}< ||\underline{{\bf x}}_d||\leq R\}$ and ${\bf 1}_A(x)$ is the indicator function for the set $A$. For $t>\frac Rc,$ formula \eqref{eq:condensintrod} registers the contribution of excursions outside $\mathbb{S}_R^{d-1}$ which are brought inside the sphere $\mathbb{S}_R^{d-1}$ by circular inversion. If $p_n(\underline{{\bf x}}_d,t)$ refers to random flights with Dirichlet distributed displacements we have that (see formulas (2.10) and (2.11) in De Gregorio and Orsingher, 2012) \begin{equation}\label{eq:condensdir} p_n(\underline{{\bf x}}_d,t)=\frac{\Gamma(\frac{n+1}{2}(d-h)+\frac h2)}{\Gamma(\frac n2(d-h))}(c^2t^2-||\underline{{\bf x}}_d||^2)^{\frac n2(d-h)-1},\quad ||\underline{{\bf x}}_d||<ct,h=1,2. \end{equation} The space-dependent factor appearing in \eqref{eq:condensdir} is a function satisfying the Euler-Poisson-Darboux (EPD) equation. In other words, the function $$f_\beta(\underline{{\bf x}}_d,t):=(c^2t^2-||\underline{{\bf x}}_d||^2)^\beta,\quad \beta\in\mathbb{R},||\underline{{\bf x}}_d||<ct,$$ satisfies the EPD equation \begin{equation} \frac{\partial ^2u}{\partial t^2}=c^2\Delta u+\frac{2\beta-1+d}{t}\frac{\partial u}{\partial t}, \end{equation} where $\Delta:=\sum_{k=1}^d\frac{\partial^2}{\partial x_k^2}$, which becomes the $d$-dimensional wave equation for $\beta=\frac{d-1}{2}$. In the reflected motions considered here we must examine the functions $$\bar f_\beta(\underline{{\bf x}}_d,t):=\left(c^2t^2-\frac{R^4}{||\underline{{\bf x}}_d||^2}\right)^\beta,\quad \beta\in\mathbb{R},\frac{R^2}{ct}<||\underline{{\bf x}}_d||<R,$$ which also are solutions of more complicated Euler-Poisson-Darboux equations where space-varying coefficients appear. The final section of the paper is concerned with random flights reflecting when colliding with hyperplanes. In this case reflection is intended in the sense that striking and reflecting paths form the same angle with respect to the normal to the hyperplane $H(\underline a_d,b)=\{\underline{{\bf x}}_d\in\mathbb{R} ^d:\, \langle\underline a_d,\underline{{\bf x}}_d\rangle=b;\,\underline a_d\in\mathbb{R}^d,b\in\mathbb{R}\}$. In this sense reflecting random flights behave as light rays of optics when the reflecting surface is a hyperplane. The trajectories of the reflecting random motions can be obtained form those of the free random flights by means of the bijective operator $\nu:\mathbb{R}^d\to \mathbb{R}^d$ defined as \begin{equation} \nu(\underline{{\bf x}}_d):=\underline{{\bf x}}_d+2\frac{ b-\langle\underline a_d,\underline{{\bf x}}_d\rangle}{\langle\underline a_d,\underline a_d\rangle}\underline a_d, \end{equation} which represents the reflection on the hyperplane $H(\underline a_d,b)$. For $t>t':=\inf(t:H(\underline a_d,b)\cap B_{ct}^d\neq \varnothing ),$ we therefore have that the density $p_n'(\underline{{\bf x}}_d,t)$ of the reflected random motion $\{\underline{\bf X}_d'(t),t>0\}$ reads \begin{align} p_{n}'(\underline{{\bf x}}_d,t) &=\begin{cases} p_n(\underline{{\bf x}}_d,t){\bf 1}_{B_{ct}^d}(\underline{{\bf x}}_d),& t\leq t'\\ p_n(\underline{{\bf x}}_d,t){\bf 1}_{L_{ct}^d}(\underline{{\bf x}}_d)+p_n\left(\nu(\underline{{\bf x}}_d),t\right){\bf 1}_{V_{ct}^d}(\underline{{\bf x}}_d),& t>t', \end{cases} \end{align} where $L_{ct}^d:=L_{ct}^d(\underline a_d, b):=\{\underline{{\bf x}}_d\in \mathbb{R}^d:||\underline{{\bf x}}_d||^2< c^2t^2,\langle\underline a_d,\underline{{\bf x}}_d\rangle<b\}$ and $V_{ct}^d$ consists of points of $B_{ct}^d$ reflected by the operator $\nu$ (see Figure \ref{fig2}). The main concepts of the inversive geometry represent fundamental tools for our analysis. Therefore, for the convenience of the reader, we recall some basic definitions in the Appendices \ref{refspheres}-\ref{refhyper}. \section{Notations} We list the main symbols used in this paper. \begin{itemize} \item With $||\cdot||$ and $\langle\cdot,\cdot\rangle$ we indicate the Euclidean norm and the scalar product, respectively. \item Let $\mathbb{S}_R^{d-1}(\underline{{\bf x}}_d^0):=\{\underline{{\bf x}}_d\in\mathbb R^d:||\underline{{\bf x}}_d-\underline{{\bf x}}_d^0||= R\}$ be the sphere with radius $R>0$ and center at $\underline{{\bf x}}_d^0\in\mathbb{R} ^d$. We set $\mathbb{S}_R^{d-1}(O):=\mathbb{S}_R^{d-1}$ where $O$ is the origin of $\mathbb{R}^d$ . Furthermore, the surface area of $\mathbb{S}_{1 }^{d-1}$ is given by $\text{area}(\mathbb{S}_{1 }^{d-1})=\frac{2\pi^{\frac d2}}{\Gamma(\frac d2)}.$ \item Let $B_R^d:=\{\underline{{\bf x}}_d\in\mathbb R^d:||\underline{{\bf x}}_d||< R\}$ be the ball with center $O$ and radius $R$. Let $C_{R_1,R_2}^d:=\{\underline{{\bf x}}_d\in\mathbb R^d:R_1< ||\underline{{\bf x}}_d||\leq R_2\}$. \item Let $H(\underline a_d,b):=\{\underline{{\bf x}}_d\in\mathbb{R} ^d:\, \langle\underline a_d,\underline{{\bf x}}_d\rangle=b;\,\underline a_d\in\mathbb{R}^d,b\in\mathbb{R}\}$ be a hyperplane in $\mathbb{R}^d$. \item Let ${\bf 1}_A(x)$ be the indicator function, that is $${\bf 1}_A(x)=\begin{cases}1,& x\in A,\\ 0,& \text{otherwise}, \end{cases}$$ while $$J_\nu(x)=\sum_{k=0}^\infty(-1)^k\left(\frac{x}{2}\right)^{2k+\nu}\frac{1}{k!\Gamma(k+\nu+1)},\quad x,\nu\in\mathbb{R},$$ is the Bessel function of order $\nu$. \item Let $n$ be the fixed number of changes of direction in the time interval $[0,t]$, we indicate the conditional probability areaure by $$P_n\{\cdot\in A\}:=P\{\cdot\in A|\mathcal N(t)=n\},$$ for all Borel sets $A$, with $n\geq 1$. Furthermore, let $E_n\{\cdot\}:=E\{\cdot|\mathcal N(t)=n\}$. \item Let $$p_n(\underline{{\bf x}}_d,t):=P_n\{\underline{{\bf X}}_d(t)\in \de\underline{{\bf x}}_d\}/\prod_{k=1}^d\de x_k,$$ $$p_n^*(\underline{{\bf x}}_d,t):=P_n\{\underline{{\bf X}}_d^*(t)\in \de\underline{{\bf x}}_d\}/\prod_{k=1}^d\de x_k,$$ $$p_n'(\underline{{\bf x}}_d,t):=P_n\{\underline{{\bf X}}_d'(t)\in \de\underline{{\bf x}}_d\}/\prod_{k=1}^d\de x_k.$$ \end{itemize} \section{Preliminary results on random flights}\label{sec:prelim} Let us indicate by $g({\bf \underline\tau}_n; t)$ the probability density function of the random vector ${\bf \underline\tau}_n.$ We provide the probability distribution of $\{\underline{{\bf X}}_d(t),t>0\}$ when the number of steps performed by the motion in $[0,t]$ is fixed. \begin{lemma} \label{lemma1} Let $n\geq 1$ be the number of changes of direction happening during time interval $[0,t]$. The conditional density function of $\{\underline{{\bf X}}_d(t),t>0\}$ is equal to \begin{align}\label{eq:densfree} p_n(\underline{{\bf x}}_d,t)=\frac{\left\{2^{\frac d2-1}\Gamma\left(\frac d2\right)\right\}^{n+1}}{(2\pi)^{\frac d2}||\underline{\bf x}_d||^{\frac d2-1}}\int_0^{\infty}\left[\int_{S_n}g({\bf \underline\tau}_n; t)\prod_{k=1}^{n+1}\left\{\frac{J_{\frac d2-1}(c\tau_k\rho)}{(c\tau_k\rho)^{\frac d2-1}} \right\}\prod_{k=1}^{n}\de\tau_k\right] \rho^{\frac d2}J_{\frac d2-1}(\rho ||\underline{\bf x}_d||) \de\rho, \end{align} with $ ||\underline{\bf x}_d||<ct$. \end{lemma} \begin{proof} The expression \eqref{eq:densfree} has been obtained by Orsingher and De Gregorio (2007) (formula (2.13)) in the uniform case and by De Gregorio (2014) (formula (3.1)) in the case of generalized Dirichlet distributions. Similar steps can be used in a general framework for the distribution $g({\bf \underline\tau}_n; t)$. In what follows, we will provide a sketch of the proof. Let us start the proof by showing that the characteristic function of $\underline{\bf X}_d(t)$ is equal to \begin{align}\label{eq:cf} \mathcal F_n(\underline{\bf \alpha}_d) &:=E_n\left\{e^{i\langle\underline{\bf \alpha}_d,\underline{{\bf X}}_d(t)\rangle}\right\}\notag\\ &=\left\{2^{\frac d2-1}\Gamma\left(\frac d2\right)\right\}^{n+1}\int_{S_n}g({\bf \underline\tau}_n; t)\prod_{k=1}^{n+1}\left\{\frac{J_{\frac d2-1}(c\tau_k||\underline{\alpha}_d||)}{(c\tau_k||\underline{\alpha}_d||)^{\frac d2-1}} \right\}\prod_{k=1}^{n}\de\tau_k. \end{align} We can write that \begin{align*} \mathcal F_n(\underline{\bf \alpha}_d)&=\int_{S_n}g({\bf \underline\tau}_n; t)\mathcal{I}_n(\underline{\alpha}_d;{\bf \underline\tau}_d)\prod_{k=1}^n\de\tau_k \end{align*} where \begin{align}\label{eq:I_n} \mathcal{I}_n(\underline{\alpha}_d;{\bf \underline\tau}_d)&:= \prod_{k=1}^{n+1}\left\{\int_{\Lambda}\exp\left\{ic\tau_k<\underline{\alpha}_d,{\bf \underline{V}}_k> \right\}\varphi(\underline{\theta}_{d-1}^k)\prod_{j=1}^{d-2}\de\theta_ j^k\de\phi^k\right\} \end{align} where $\Lambda:=[0,\pi]^n\times[0,2\pi]$. It is well-known that the integral $\mathcal{I}_n(\underline{\alpha}_d;{\bf \underline\tau}_d)$ (see Theorem 2.1 of Orsingher and De Gregorio, 2007, and formula (2.5) of De Gregorio and Orsingher, 2012) is equal to \begin{equation}\label{intangle} \mathcal{I}_n(\underline{\alpha}_d;{\bf \underline\tau}_d)=\left\{2^{\frac d2-1}\Gamma\left(\frac d2\right)\right\}^{n+1}\prod_{k=1}^{n+1}\frac{J_{\frac d2-1}(c\tau_k||\underline{\alpha}_d||)}{(c\tau_k||\underline{\alpha}_d||)^{\frac d2-1}} \end{equation} and this leads to \eqref{eq:cf}. Now, by inverting the characteristic function \eqref{eq:cf}, we are able to show that the density of the process $\{\underline{\bf X}_d(t),t>0\},$ is given by \eqref{eq:densfree}. Let us denote by $\underline{v}_d$ the vector $$\underline{v}_d=\left( \begin{array}{l} \sin\theta_{1}\sin\theta_{2}\cdot\cdot\cdot\sin\theta_{d-2}\sin\phi \\ \sin\theta_{1}\sin\theta_{2}\cdot\cdot\cdot\sin\theta_{d-2}\cos\phi \\ ...\\ \sin\theta_{1}\cos\theta_{2}\\ \cos\theta_{1} \end{array} \right).$$ Therefore, by inverting the characteristic function \eqref{eq:cf} and by passing to $d$-dimensional spherical coordinates, we have, for $\underline{{\bf x}}_d\in B_{ct}^d$, that \begin{align*} &p_n(\underline{\bf x}_d,t)\\ &=\frac{1}{(2\pi)^d}\int_{\mathbb{R}^d}e^{-i<\underline{\alpha}_d,\underline{\bf x}_d>}\mathcal F_n(\underline{\bf \alpha}_d)\prod_{k=1}^d\de\alpha_k\notag\\ &=\frac{1}{(2\pi)^d}\int_0^\infty \rho^{d-1}\de\rho\int_{\Lambda} e^{-i\rho<\underline{v}_d,\underline{\bf x}_d>}\de\mathbb {S}_1^{d-1}\left\{2^{\frac d2-1}\Gamma\left(\frac d2\right)\right\}^{n+1}\int_{S_n}g({\bf \underline\tau}_n; t)\prod_{k=1}^{n+1}\left\{\frac{J_{\frac d2-1}(c\tau_k\rho)}{(c\tau_k\rho)^{\frac d2-1}} \right\}\prod_{k=1}^{n}\de\tau_k. \end{align*} where $\de\mathbb{S}_{1 }^{d-1}:=\prod_{i=1}^{d-2}\{\sin^{d-i-1}\theta_i\de\theta_i\}\de\phi.$ By applying formula (2.15) of Orsingher and De Gregorio (2007) \begin{equation}\label{eq: (2.12) DGO12} \int_{\Lambda} e^{-i\rho<\underline{v}_d,\underline{\bf x}_d>}\de\mathbb {S}_1^{d-1}=(2\pi)^{\frac d2}\frac{J_{\frac d2-1}(\rho ||\underline{\bf x}_d||)}{(\rho ||\underline{\bf x}_d||)^{\frac d2-1}}, \end{equation} we arrive at the claimed result. \end{proof} In the analysis of random flights, a central role is played by the probabilistic assumptions on the displacements $c{\bf \underline\tau}_n$. It is useful to assume that the random vector ${\bf \underline\tau}_n$ has the following Dirichlet distribution \begin{equation}\label{eq:dirichlet} g({\bf \underline\tau}_n;\underline{\bf q}_{n+1},t)=\frac{\Gamma\left((n+1)\left(\frac dh-1\right)\right)}{(\Gamma\left(\frac dh-1\right))^{n+1}}\frac{1}{t^{(n+1)\left(\frac dh-1\right)}}\prod_{k=1}^{n+1}\tau_k^{\frac dh-2}, \end{equation} where ${\bf \underline\tau}_n\in S_n$, with parameters $\underline{\bf q}_{n+1}:=(\frac dh-1,...,\frac dh-1),h:=1,2,$ and the conditions $d\geq 2$ if $h=1$ and $d\geq 3$ if $h=2$ hold. Under the assumptions \eqref{eq:dirichlet}, the density function \eqref{eq:densfree} can be evaluated explicitly \begin{equation}\label{eq:conddensfree} p_{n,h}(\underline{{\bf x}}_d,t)=\frac{\Gamma(\frac{n+1}{2}(d-h)+\frac h2)}{\Gamma(\frac n2(d-h))}\frac{(c^2t^2-||\underline{{\bf x}}_d||)^{\frac n2(d-h)-1}}{\pi^{\frac d2}(ct)^{(n+1)(d-h)+h-2}}{\bf 1}_{B_R^d}(\underline{{\bf x}}_d), \end{equation} while for the process $\{D_d(t)=||\underline{\bf X}_d(t)||,t>0\},$ we are able to obtain the following conditional density \begin{equation}\label{eq:densradialfree} q_{n,d,h}(r,t)=\frac{2\Gamma(\frac{n+1}{2}(d-h)+\frac h2)}{\Gamma(\frac d2)\Gamma(\frac n2(d-h))}\frac{r^{d-1}(c^2t^2-r^2)^{\frac n2(d-h)-1}}{(ct)^{(n+1)(d-h)+h-2}}{\bf 1}_{(0,R)}(r), \end{equation} (see De Gregorio and Orsingher (2012) and Le Ca\"er (2010)). We observe that $\{D_d(t),t>0\},$ is related to the Beta distribution. Indeed, $D_d(t)/c^2t^2$ is distributed as a Beta random variable with parameters $\frac d2$ and $\frac n2(d-h)$. \section{Reflecting random flights in spheres} \subsection{Definition and probability distributions} Let us consider a random flight $\{\underline{{\bf X}}_d(t),t>0\},$ defined as in the Introduction. When a sample path of $\underline{{\bf X}}_d(t)$ strikes the sphere $\mathbb{S}^{d-1}_R$, the trajectory of the process is reflected inside $\mathbb{S}^{d-1}_R$. The reflection of the random flight on the boundary $\mathbb{S}_R^{d-1}$ can be envisaged in different ways. The specular reflection (the incoming sample path forms the same angle as the reflected trajectory with respect to the normal vector to $\mathbb{S}_R^{d-1}$) seems the most natural one. Nevertheless, from the mathematical point of view, the reflecting surface is closed and this implies that the probability distribution of the reflected process takes a cumbersome form for sufficiently large values of $t$. For this reason, we assume that the reflection is based on the principle of circular inversion in spheres defined in Appendix \ref{refspheres}. The most important effect of this procedure is that the sample paths obtained by reflection are deformed and take the structure of circumference arcs. This leads to a new process, namely, the reflecting random flight moving in $B^{d}_R\cup \mathbb{S}^{d-1}_R$. \begin{definition}\label{def:refsp} The reflecting random flight $\{\underline{{\bf X}}_d^*(t),t>0\},$ reflected on the sphere $\mathbb{S}_R^{d-1}$ is constructed by means of the free process $\{\underline{{\bf X}}_d(t),t>0\}$ as follows: 1) if $t\leq \frac Rc,$ then $\underline{{\bf X}}_d^*(t) :=\underline{{\bf X}}_d(t)$; 2) if $t>\frac Rc,$ then if at least one change of direction happens during the time interval $[0,t]$, we have that \begin{align}\label{eq:def} \underline{{\bf X}}_d^*(t) :=\underline{{\bf X}}_d(t){\bf 1}_{B_R^d}(\underline{{\bf X}}_d(t))+\mu_R(\underline{{\bf X}}_d(t) {\bf 1}_{C^d_{R,ct}}(\underline{{\bf X}}_d(t)), \end{align} where $\mu_R(\underline{{\bf x}}_d)=R^2\frac{\underline{{\bf x}}_d}{||\underline{{\bf x}}_d||^2}$ is the inversion map defined by \eqref{eq:appendixmap1}; while if there are no deviations \begin{align}\label{eq:def2} \underline{{\bf X}}_d^*(t) :=\mu_R(\underline{{\bf X}}_d(t) {\bf 1}_{\mathbb{S}^{d-1}_{ct}}(\underline{{\bf X}}_d(t)). \end{align} \end{definition} Definition \ref{def:refsp} leads to the following considerations on the sample paths of $\{\underline{{\bf X}}_d^*(t),t>0\}$: \begin{itemize} \item the reflecting random flight has two components: the first one is given by the free process $\{\underline{{\bf X}}_d(t),t>0\}$; the second component $\mu_R(\underline{{\bf X}}_d(t))$ is due to the reflection in $\mathbb{S}_R^{d-1}$ of the sample paths of $\{\underline{{\bf X}}_d(t),t>0\},$ wandering outside the sphere $\mathbb{S}_R^{d-1}$. It is worth mentioning that the property P4) implies that the reflected paths have opposite orientation w.r.t. the trajectories of $\{\underline{{\bf X}}_d(t),t>0\}$ moving outside $\mathbb{S}_R^{d-1}$; \item the reflecting component of $\{\underline{{\bf X}}_d^*(t),t>0\}$ is given by \begin{align}\label{eq:refcomp} \mu_R(\underline{{\bf X}}_d(t))=\frac{R^2}{||\underline{{\bf X}}_d(t)||^2}\underline{{\bf X}}_d(t)=\frac{cR^2}{||\underline{{\bf X}}_d(t)||^2}\sum_{k=1}^{n+1}{\bf \underline{V}}_k\tau_k, \end{align} for $\underline{{\bf X}}_d(t)\in C^d_{R,ct}$. Therefore, $\frac{cR^2}{||\underline{{\bf X}}_d(t)||^2}$ can be thought of as the random velocity of \eqref{eq:refcomp}. Since $\frac{cR^2}{||\underline{{\bf X}}_d(t)||^2}\leq c$, the reflected motion travels more slowly than the one related to the process $\{\underline{{\bf X}}_d(t),t>0\}$. For instance, see sample path ${\bf b}$ in Figure \ref{fig1}; \item the property P5) of the map $\mu_R$ implies that the sample paths of $\{\underline{{\bf X}}_d^*(t),t>0\}$ can be represented by broken lines, when no reflection has occurred, and by the composition of straight lines and circumference arcs, when at least one reflection has taken place (see sample path a in Figure \ref{fig1}). \end{itemize} \begin{figure}[t] \begin{center} \includegraphics[angle=0,width=0.7\textwidth]{reflpathsext.eps} \caption{Two typical sample paths are depicted. The trajectory ${\bf a}$ is composed of straight lines and arc of spheres, while the sample path ${\bf b}$ is obtained if no changes of direction happen in the interval $[0,t]$.}\label{fig1} \end{center} \end{figure} We are able to provide the conditional probability distributions $\{\underline{{\bf X}}_d^*(t),t>0\},$ by means of those of $\{\underline{{\bf X}}_d(t),t>0\}$. Therefore, by means of \eqref{eq:densfree}, we are able to obtain the next result. \begin{theorem} If $\mathcal N(t)=n$, with $n\geq 1,$ the process $\{\underline{{\bf X}}_d^*(t),t>0\},$ has the following conditional density function \begin{align}\label{eq:condens} p_{n}^*(\underline{{\bf x}}_d,t)=\begin{cases} p_n(\underline{{\bf x}}_d,t){\bf 1}_{B_R^d}(\underline{{\bf x}}_d),& t\leq\frac{R}{c}\\ p_n(\underline{{\bf x}}_d,t){\bf 1}_{B_R^d}(\underline{{\bf x}}_d)+\frac{R^{2d}}{||\underline{{\bf x}}_d||^{2d}}p_n\left(R^2\frac{\underline{{\bf x}}_d}{||\underline{{\bf x}}_d||^2},t\right){\bf 1}_{C^d_{\frac{R^2}{ct},R}}(\underline{{\bf x}}_d),& t>\frac{R}{c}. \end{cases} \end{align} where $p_n(\underline{{\bf x}}_d,t)$ is equal to \eqref{eq:densfree}. \end{theorem} \begin{proof} The case $ t\leq\frac{R}{c}$ is obvious. Now we assume that $t>\frac Rc$. Let $A$ be a Borel set such that $A\cap \overline{B_{R}^d}\neq \varnothing$. Let $\underline{{\bf Y}}_d(t):=\mu_R(\underline{{\bf X}}_d(t))=R^2\frac{\underline{{\bf X}}_d(t)}{||\underline{{\bf X}}_d(t)||^2},$ we get \begin{align}\label{eq:step1dens} P_n\{\underline{{\bf X}}_d^*(t)\in A\}&=P_n\{\underline{{\bf X}}_d(t)\in A\cap B_{R}^d\}+P_n\left\{\underline{{\bf Y}}_d(t)\in A\cap C_{\frac{R^2}{ct},R}^d\right\}\notag\\ &=\int_{ A\cap B_{R}^d}p_n(\underline{{\bf x}}_d,t)\prod_{k=1}^d\de x_k+P_n\left\{\underline{{\bf Y}}_d(t)\in A\cap C^d_{\frac{R^2}{ct},R}\right\}. \end{align} The first term in the previous expression refers to the sample paths which do not attain the boundary $\mathbb{S}_R^{d-1}$ up to time $t,$ while the second probability concerns the trajectories reflecting on $\mathbb{S}_R^{d-1}$. By setting $$D_A:=\mu_R^{-1}\left(A\cap C_{\frac{R^2}{ct},R}^d\right)=\left\{\underline{{\bf y}}_d\in \mathbb{R}^d: \underline{{\bf x}}_d=\mu_R(\underline{{\bf y}}_d)\in A\cap C^d_{\frac{R^2}{ct},R}\right\}$$ then, by means of Jacobi transformation formula, we have that \begin{align}\label{eq:step2dens} P_n\left\{\underline{{\bf Y}}_d(t)\in A\cap C^d_{\frac{R^2}{ct},R}\right\}&=P_n\{\underline{{\bf X}}_d(t)\in D_A\}\notag\\ &=\int_{D_A}p_n(\underline{{\bf y}}_d,t)\prod_{k=1}^d\de y_k\notag\\ &=\int_{A\cap C^d_{\frac{R^2}{ct},R}} p_n\left(\mu_R^{-1}(\underline{{\bf x}}_d),t\right)|\text{det}(J_{\mu_R^{-1}}(\underline{{\bf x}}_d))|\prod_{k=1}^d \de x_k\notag\\ &=\int_{A\cap C^d_{\frac{R^2}{ct},R}} p_n\left(\mu_R(\underline{{\bf x}}_d),t\right)|\text{det}(J_{\mu_R}(\underline{{\bf x}}_d))|\prod_{k=1}^d \de x_k \end{align} where in the last step we have used the following facts: $\mu_R=\mu_R^{-1}$ (which follows from the property P3)) and $|\text{det}(J_{\mu_R}(\underline{{\bf x}}_d))|=\frac{R^{2d}}{||\underline{{\bf x}}_d||^{2d}}$. From \eqref{eq:step1dens} and \eqref{eq:step2dens}, the result \eqref{eq:condens} immediately follows. \end{proof} From Lemma \ref{lemma1} emerges that the random process $\{\underline{\bf X}_d(t),t>0\}$ is isotropic, namely $Q(\underline{\bf X}_d(t))\sim \underline{\bf X}_d(t)$ for all $Q\in O(d)$ and $p_n(\underline{{\bf x}}_d,t)=p_n(||\underline{{\bf x}}_d||,t)$. Furthermore, from \eqref{eq:densfree}, we have that $\{\underline{\bf X}_d^*(t),t>0\}$ is invariant by rotation as well. Therefore, $p_n^*(\underline{{\bf x}}_d,t)$ depends on the distance $||\underline{{\bf x}}_d||$ and thus we can write \begin{align}\label{eq:rotinv} p_n^*(\underline{{\bf x}}_d,t)&=p_n^*(||\underline{{\bf x}}_d||,t)\notag\\ &=p_n(||\underline{{\bf x}}_d||,t){\bf 1}_{(0,R)}(||\underline{{\bf x}}_d||)+\frac{R^{2d}}{||\underline{{\bf x}}_d||^{2d}}p_n\left(\frac{R^2}{||\underline{{\bf x}}_d||},t\right){\bf 1}_{(R^2/ct,R]}(||\underline{{\bf x}}_d||) \end{align} for $t>\frac Rc.$ \begin{definition} The reflecting radial process $\{D_d^*(t),t>0\},$ represents the Euclidean distance from $\underline{{\bf 0}}_d$ of the position $\underline{{\bf X}}_d^*(t)$, namely $D_d^*(t)=||\underline{{\bf X}}_d^*(t)||$. It can be defined by \begin{align} D_d^*(t)=D_d(t){\bf 1}_{(0,R)}(D_d(t))+\frac{R^2}{D_d(t)}{\bf 1}_{[R,ct]}(D_d(t)), \end{align} where $D_d(t) =||\underline{{\bf X}}_d(t)||$. \end{definition} As a consequence of the isotropic structure of $\{\underline{\bf X}_d^*(t),t>0\}$, the conditional density function of $\{D_d^*(t),t>0\},$ becomes \begin{equation}\label{eq:distancecircinv} q_{n,d}^*(r,t)=r^{d-1}\text{area}(\mathbb{S}_{1}^{d-1})p_n^*(r,t){\bf 1}_{(0,R]}(r). \end{equation} \subsection{Reflecting Dirichlet random flights} A suitable choice of the distribution $g({\bf \underline\tau}_n; t)$ leads to explicit expressions for the conditional density functions \eqref{eq:condens}. As we have seen in Section \ref{sec:prelim} the Dirichlet distributions play a special role in the study of random flights. The assumption \eqref{eq:dirichlet} and the results \eqref{eq:conddensfree} and \eqref{eq:densradialfree} imply that the reflecting process $\{\underline{{\bf X}}_d^*(t),t>0\}$ has probability law \eqref{eq:condens} given by \begin{align}\label{eq:densrefl} p_{n,h}^*(\underline{{\bf x}}_d,t)&=\frac{\Gamma(\frac{n+1}{2}(d-h)+\frac h2)}{\Gamma(\frac n2(d-h))}\frac{1}{\pi^{\frac d2}(ct)^{(n+1)(d-h)+h-2}}\\ &\quad\times\left[(c^2t^2-||\underline{{\bf x}}_d||^2)^{\frac n2(d-h)-1}{\bf 1}_{B_R^d}(\underline{{\bf x}}_d)+\frac{R^{2d}}{||\underline{{\bf x}}_d||^{2d}}\left(c^2t^2-\frac{R^4}{||\underline{{\bf x}}_d||^2}\right)^{\frac n2(d-h)-1}{\bf 1}_{C^d_{\frac{R^2}{ct},R}}(\underline{{\bf x}}_d)\right],\notag \end{align} with $t>\frac Rc$, while $\{D_d^*(t),t>0\}$ has the following conditional distribution \begin{align}\label{eq:radialcondensdir} q_{n,d,h}^*(r,t)&=\frac{2\Gamma(\frac{n+1}{2}(d-h)+\frac h2)}{\Gamma(\frac d2)\Gamma(\frac n2(d-h))(ct)^{(n+1)(d-h)+h-2}}\\ &\quad\times\left[r^{d-1}(c^2t^2-r^2)^{\frac n2(d-h)-1}{\bf 1}_{(0,R)}(r)+\frac{R^{2d}}{r^{d+1}}\left(c^2t^2-\frac{R^4}{r^2}\right)^{\frac n2(d-h)-1}{\bf 1}_{(R^2/ct,R]}(r)\right],\notag \end{align} with $t>\frac Rc$. In this case we call these random walks ``reflecting Dirichlet random flights''. \begin{remark}\label{remarkun} The probability $P_n\{\underline{{\bf X}}_d(t)\in \de \underline{{\bf x}}_d\}$ is uniform inside the ball $B_{ct}^d$ in the following cases: (i) $h=1, d=2, n=2$; (ii) $h=1, d=3, n=1$; (iii) $h=2, d=3, n=2$; (iv) $h=2, d=4, n=1$. In the cases (i)-(iv), we have that the function $p_{n,h}^*(\underline{{\bf x}}_d,t)$ becomes \begin{align*} p_{2,1}^*(\underline{{\bf x}}_2,t)&=\frac{1}{\pi(ct)^2}\left[{\bf 1}_{B_R^2}(\underline{{\bf x}}_2)+\frac{R^{4}}{||\underline{{\bf x}}_2||^{4}}{\bf 1}_{C^2_{\frac{R^2}{ct},R}}(\underline{{\bf x}}_2)\right], \\ p_{1,1}^*(\underline{{\bf x}}_3,t)&=p_{2,2}^*(\underline{{\bf x}}_3,t)=\frac{\Gamma(\frac 52)}{\pi^{\frac 32}(ct)^3}\left[{\bf 1}_{B_R^3}(\underline{{\bf x}}_3)+\frac{R^{6}}{||\underline{{\bf x}}_3||^{6}}{\bf 1}_{C^3_{\frac{R^2}{ct},R}}(\underline{{\bf x}}_3)\right], \\ p_{1,2}^*(\underline{{\bf x}}_4,t)&=\frac{\Gamma(3)}{\pi^2(ct)^4}\left[{\bf 1}_{B_R^4}(\underline{{\bf x}}_4)+\frac{R^{8}}{||\underline{{\bf x}}_4||^{8}}{\bf 1}_{C^4_{\frac{R^2}{ct},R}}(\underline{{\bf x}}_4)\right]. \end{align*} This implies that the reflecting random flights are never uniformly distributed inside the ball $B_{R}^d$. \end{remark} Now, we focus our attention on the distribution function of $\{D_d^*(t),t>0\}$ which also provides information on the probability of the position of the reflecting random flight at time $t>0$, that is $$P_n\{D_d^*(t)<r\}=P_n\{\underline{\bf X}_d^*(t)\in B_r^d\}, \quad 0<r\leq R.$$ The probability distribution of the distance process $\{D_d^*(t),t>0\}$ is related to the Beta distribution as shown in the next result. Let $r\in(0,R]$ and $t>\frac Rc$, by using \eqref{eq:densradialfree}, we immediately obtain that \begin{align}\label{eq:probball2} P_n\{ D_d^*(t)<r\}&=\int_0^{\frac{r^2}{(ct)^2}}\frac{x^{\frac d2-1}(1-x)^{\frac n2(d-h)-1}}{B\left(\frac d2,\frac n2(d-h)\right)}\de x+\left[\int_{\frac{R^4}{(ctr)^2}}^1\frac{x^{\frac d2-1}(1-x)^{\frac n2(d-h)-1}}{B\left(\frac d2,\frac n2(d-h)\right)}\de x\right]{\bf 1}_{(R^2/ct,R]}(r)\notag\\ &=\frac{B\left(\frac{r^2}{(ct)^2};\frac d2,\frac n2(d-h)\right)}{B\left(\frac d2,\frac n2(d-h)\right)}+\left[1-\frac{B\left(\frac{R^4}{(ctr)^2};\frac d2,\frac n2(d-h)\right)}{B\left(\frac d2,\frac n2(d-h)\right)}\right]{\bf 1}_{(R^2/ct,R]}(r) \end{align} where $B(a,b):=\Gamma(a)\Gamma(b)/\Gamma(a+b)$ and $B(z;a,b):=\int_0^zu^{a-1}(1-u)^{b-1}du$ is the incomplete gamma function. From \eqref{eq:probball2} we obtain that \begin{align*} P_n\{\underline{\bf X}_d^*(t)\in C^d_{\frac{R^2}{ct},R}\}&=P_n\left\{\frac{R^2}{ct}<D_d^*(t)\leq R\right\}\\ &=1-B\left(R^4;\frac d2,\frac n2(d-h)\right) \end{align*} The probability \eqref{eq:probball2} becomes particularly simple for $d=2$ and $h=1$. Indeed, we have that \begin{align} P_n\{ D_2^*(t)<r\}&=1-\left(1-\frac{r^2}{(ct)^2}\right)^{\frac n2}+\left(1-\frac{R^4}{(ctr)^2}\right)^{\frac n2}{\bf 1}_{(R^2/ct,R]}(r),\quad r\in(0,R]. \end{align} Furthermore, if we assume that $h=2, d=2d',d'\geq 2,$ it is possible to write down $P_n\{ D_d^*(t)<r\}$ by means of the probability distribution of binomial r.v.'s. By exploiting the following well-known result \begin{equation}\label{eq:relincombeta} \frac{B(x;a,b)}{B(a,b)}=\sum_{k=a}^{a+b-1}\binom{a+b-1}{k}x^k(1-x)^{a+b-1-k},\quad a,b\in\mathbb N, \end{equation} it is not hard to prove that \begin{align} P_n\{D_d^*(t)<r\}&=P\left\{\frac d2\leq Y_{n,d}\leq\frac{n+1}{2}(d-2)\right\}+P\left\{0\leq \hat Y_{n,d}<\frac{d}{2}\right\}{\bf 1}_{(R^2/ct,R]}(r) \end{align} where $Y_{n,d}\sim$Bin$\left(\frac{n+1}{2}(d-2),\frac{r^2}{(ct)^2}\right)$ and $\hat Y_{n,d}\sim$Bin$\left(\frac{n+1}{2}(d-2),\frac{R^4}{(ctr)^2}\right)$. For $t>\frac Rc$, we are also able to derive the $m$-th moment, with $m\geq 1,$ of $\{D_d^*(t),t>0\}$. We have that \begin{align}\label{eq:moments} E_n\{D_d^*(t)\}^m&=\frac{1}{B\left(\frac d2,\frac n2(d-h)\right)}\left[(ct)^m\int_0^{\frac{R^2}{(ct)^2}}x^{\frac {d+m}{2}-1}(1-x)^{\frac{n}{2}(d-h)-1}\de x\right.\notag\\ &\quad\left.+\left(\frac{R^2}{ct}\right)^{m}\int_{\frac{R^2}{(ct)^2}}^1x^{\frac {d-m}{2}-1}(1-x)^{\frac{n}{2}(d-h)-1}\de x\right]. \end{align} Moreover, if $d>m,$ we can write \eqref{eq:moments} in terms of beta and incomplete beta functions as follows \begin{align}\label{eq:moments2} E_n\{D_d^*(t)\}^m &=(ct)^m\frac{B\left(\frac{R^2}{(ct)^2};\frac {d+m}{2},\frac n2(d-h)\right)}{B\left(\frac {d}{2},\frac n2(d-h)\right)}+\left(\frac{R^2}{ct}\right)^{m}\left[1-\frac{B\left(\frac{R^2}{(ct)^2};\frac{ d-m}{2},\frac n2(d-h)\right)}{B\left(\frac d2,\frac n2(d-h)\right)}\right].\end{align} \subsection{Random flights and the Euler-Poisson-Darboux equation} In De Gregorio and Orsingher (2012) (Remark 2.7) is observed that the functions \begin{equation}\label{eq:epd} f_\beta(\underline{{\bf x}}_d,t):=(c^2t^2-||\underline{{\bf x}}_d||^2)^\beta,\quad ||\underline{{\bf x}}_d||<ct, \beta\in\mathbb{R}, \end{equation} satisfy the following Euler-Poisson-Darboux (EPD) partial differential equation \begin{equation}\label{eq:epdcon} \frac{\partial ^2u}{\partial t^2}-\frac{2\beta-1+d}{t}\frac{\partial u}{\partial t}=c^2\Delta u \end{equation} where $\Delta:=\sum_{k=1}^d\frac{\partial^2}{\partial x_k^2}$. If $\beta=\frac{n}{2}(d-h)-1,$ we have that $$f_{\frac{n}{2}(d-h)-1}(\underline{{\bf x}}_d,t)=\frac{\pi^{\frac d2}(ct)^{(n+1)(d-h)+h-2}\Gamma(\frac n2(d-h))}{\Gamma(\frac{n+1}{2}(d-h)+\frac h2)}p_{n,h}(\underline{{\bf x}}_d,t).$$ Therefore, by exploiting the equation \eqref{eq:epdcon}, it is not hard to show that $p_{n,h}(\underline{{\bf x}}_d,t)$ is a solution of the following EPD equation \begin{equation}\label{eq:epdcon2} \frac{\partial ^2u}{\partial t^2}+\frac{(n+1)(d-h)+h-1}{t}\frac{\partial u}{\partial t}=c^2\Delta u. \end{equation} The projection of the random process $\{\underline{{\bf X}}_d(t),t>0\}$ onto a lower space of dimension $m,$ implies that the conditional marginal distributions become \begin{equation*} p_{n,h}^d(\underline{{\bf x}}_m,t)=\frac{\Gamma(\frac{n+1}{2}(d-h)+\frac h2)}{\Gamma(\frac{ n+1}{2}(d-h)+\frac{h-m}{2})}\frac{(c^2t^2-||\underline{{\bf x}}_m||)^{\frac{n+1}{2}(d-h)-\frac{m-h}{2}-1}}{\pi^{\frac m2}(ct)^{(n+1)(d-h)+h-2}}, \end{equation*} (see formulas (2.26) and (2.27) of De Gregorio and Orsingher, 2012). By means of the same considerations used for the density function $p_{n,h}(\underline{{\bf x}}_d,t)$ we obtain that $p_{n,h}^d(\underline{{\bf x}}_m,t)$ is still solution of the EPD equation \eqref{eq:epdcon}. In the same spirit of the previous considerations, the function $$\bar f_\beta(\underline{{\bf x}}_d,t):=\left(c^2t^2- \frac{R^4}{||\underline{{\bf x}}_d||^2}\right)^\beta,\quad\frac{R^2}{ct}< ||\underline{{\bf x}}_d||\leq R, \beta\in\mathbb{R}, $$ solves the following EPD partial differential equation with time and space varying coefficients \begin{equation}\label{eq:epd2} \frac{\partial ^2u}{\partial t^2}-\frac{a_\beta(\underline{{\bf x}}_d)}{t}\frac{\partial u}{\partial t}=c^2\Delta u,\end{equation} where $$a_\beta(\underline{{\bf x}}_d):=2\beta-1+\frac{R^4}{||\underline{{\bf x}}_d||^2}\left[\frac{4-d}{||\underline{{\bf x}}_d||^2}+2(\beta-1)\left(\frac{1-R^4/||\underline{{\bf x}}_d||^4}{c^2t^2-R^4/||\underline{{\bf x}}_d||^2}\right)\right].$$ For $d=4$ and $\beta=1$, we have that the function $\bar f_1(\underline{{\bf x}}_4,t)$ satisfies the equation $$\frac{\partial ^2u}{\partial t^2}-\frac{1}{t}\frac{\partial u}{\partial t}=c^2\Delta u.$$ By setting $$\bar p_{n,h}(\underline{{\bf x}}_d,t):=\frac{\Gamma(\frac{n+1}{2}(d-h)+\frac h2)}{\Gamma(\frac n2(d-h))}\frac{\frac{R^{2d}}{||\underline{{\bf x}}_d||^{2d}}}{\pi^{\frac d2}(ct)^{(n+1)(d-h)+h-2}}\left(c^2t^2-\frac{R^4}{||\underline{{\bf x}}_d||^2}\right)^{\frac n2(d-h)-1},$$ with $\frac{R^2}{ct}< ||\underline{{\bf x}}_d||\leq R,$ we have that $$\bar f_{\frac{n}{2}(d-h)-1}(\underline{{\bf x}}_d,t)=\frac{\pi^{\frac d2}(ct)^{(n+1)(d-h)+h-2}\Gamma(\frac n2(d-h))}{\Gamma(\frac{n+1}{2}(d-h)+\frac h2)}\frac{||\underline{{\bf x}}_d||^{2d}}{R^{2d}}\bar p_{n,h}(\underline{{\bf x}}_d,t).$$ Therefore, by exploiting the equation \eqref{eq:epd2}, we conclude that $\bar p_{n,h}(\underline{{\bf x}}_d,t)$ is a solution of the following partial differential equation \begin{align}\label{eq:pderef} &\frac{\partial ^2u}{\partial t^2}+\frac{2(2\beta+d)-a_\beta(\underline{{\bf x}}_d)}{t}\frac{\partial u}{\partial t}+\left[\frac{(2\beta+d)((2\beta+d-1)-a_\beta(\underline{{\bf x}}_d))}{t^2}\right]u\notag\\ &=c^2\left[\Delta +\frac{2d(3d-2)}{||\underline{{\bf x}}_d||^2}+\frac{4d\langle \underline{{\bf x}}_d,\nabla \rangle}{||\underline{{\bf x}}_d||^2}\right]u, \end{align} where $\beta=\frac{n}{2}(d-h)-1$ and $\nabla u:=$grad$u$. The equation \eqref{eq:pderef} no longer has the structure of the EPD equation. \subsection{On the unconditional probability distributions} In order to obtain unconditional densities for $\{\underline{\bf X}_d^*(t),t>0\},$ we should specify the probability distribution of $\mathcal N(t)$. Different choices of the above probability law lead to different unconditional densities of the reflecting random flight. We assume that the number of deviations $\mathcal{N}(t)=\mathcal{N}_d(t), d\geq 2$, at time $t>0$, possesses the distribution of a weighted Poisson random variable. Let $\{N(t),t>0\}$ be a homogeneous Poisson process with rate $\lambda>0,$ a random variable $\mathcal{N}_d(t)$ has weighted Poisson probability distribution if \begin{equation} P\{\mathcal{N}_d(t)=n\}=\frac{w_n P\{N(t)=n\}}{\sum_{k=0}^\infty w_k P\{N(t)=k\}}, \quad n\in\mathbb N_0, \end{equation} where $w_ks$ are non-negative weight functions with $0<\sum_{k=0}^\infty w_k P\{N(t)=k\}<\infty$ (for more details see Balakrishnan and Kozubowski, 2008). In our context, we choose the weights as follows $$w_k=\frac{k!}{\Gamma((\frac{d-h}{2})k+\frac d2)}$$ and then for this choice we obtain that \begin{equation}\label{lawgenpoi} P\left\{\mathcal{N}_d(t)=n\right\}=\frac{1}{E_{\frac{d-h}{2},\frac d2}(\lambda t)}\frac{(\lambda t)^n}{\Gamma((\frac{d-h}{2})n+\frac d2)},\quad n\in\mathbb{N}_0, \end{equation} where $E_{\alpha,\beta}(x)=\sum_{k=0}^\infty \frac{x^k}{\Gamma(\alpha k+\beta)},\, x\in\mathbb{R},\alpha,\beta>0,$ is the two-parameter Mittag-Leffler function. The random variable $\mathcal{N}_d(t)$ with probability distribution \eqref{lawgenpoi} coincides with (4.1) in Beghin and Orsingher (2009). \begin{theorem} For $t>\frac Rc$ and by assuming \eqref{lawgenpoi}, we have that the unconditional probability distribution of $\{\underline{\bf X}_{d}^*(t),t>0\}$ becomes \begin{equation}\label{eq:comprdis} P\{\underline{\bf X}_{d}^*(t)\in \de\underline{\bf x}_{d}\}=p_h^*(\underline{\bf x}_d,t)\prod_{k=1}^d\de x_k+\frac{1}{E_{\frac{d-h}{2},\frac d2}(\lambda t)\Gamma(\frac d2)}\mu^*(\de\underline{\bf x}_d ) \end{equation} where \begin{align}\label{eq:unconddensrefl0} p_h^*(\underline{\bf x}_d,t)&=\frac{1}{(ct)^d\pi^{\frac d2}}\frac{1}{E_{\frac{d-h}{2},\frac d2}(\lambda t)}\left\{\left[\gamma_h(||\underline{\bf x}_d||,t)\right]^{1-\frac{2}{d-h}}E_{\frac {d-h}{2},\frac {d-h}{2}}\left(\gamma_h(||\underline{\bf x}_d||,t)\right){\bf 1}_{B_R^d}(\underline{{\bf x}}_d)\right.\notag\\ &\quad+\left.\frac{R^{2d}}{||\underline{{\bf x}}_d||^{2d}}\left[\gamma_h\left(\frac{R^2}{||\underline{\bf x}_d||},t\right)\right]^{1-\frac{2}{d-h}}E_{\frac {d-h}{2},\frac {d-h}{2}}\left(\gamma_h\left(\frac{R^2}{||\underline{\bf x}_d||},t\right)\right){\bf 1}_{C^d_{\frac{R^2}{ct},R}}(\underline{{\bf x}}_d)\right\}, \end{align} with $$\gamma_h(||\underline{\bf x}_d||,t):=\lambda t\left(1-\frac{||\underline{\bf x}_{d}||^2}{c^2t^2}\right)^{\frac {d-h}{2}},$$ while $\mu^*$ is the uniform distribution on $\mathbb{S}^{d-1}_{\frac{R^2}{ct}}.$ \end{theorem} \begin{proof} The second term in \eqref{eq:comprdis} arises from the following considerations. The probability distribution of the random flight $\{\underline{\bf X}_d^*(t),t>0\}$ admits a singular component emerging in the case $\mathcal{N}_d(t)=0,$ that is if the initial direction of the motion does not change up to time $t$. For $t>\frac Rc,$ the circular inversion of $\mathbb S_{ct}^{d-1}$ with respect to $\mathbb S_R^{d-1}$ leads to $\mathbb S_{\frac{R^2}{ct}}^{d-1}$. Therefore, if there are no changes of direction in $(0,t]$, the reflected path reaches $\mathbb S_{\frac{R^2}{ct}}^{d-1}$ and $$P\left\{\underline{\bf X}_d^*(t)\in\mathbb S_{\frac{R^2}{ct}}^{d-1}\right\}=P\{\mathcal{N}_d(t)=0\}=\frac{1}{E_{\frac{d-h}{2},\frac d2}(\lambda t)\Gamma(\frac d2)}.$$ For the remaining part of the distribution $P\{\underline{\bf X}_{d}^*(t)\in \de\underline{\bf x}_{d}\}$ we can observe that \begin{align*} p_h^*(\underline{\bf x}_d,t)&=\sum_{n=1}^\infty p_{n,h}^*(\underline{\bf x}_d,t)P\{\mathcal N_d(t)=n\} \end{align*} where $p_{n,h}(\underline{\bf x}_d,t)$ is defined by \eqref{eq:densrefl}. The same steps performed in the proof of Theorem 5 in De Gregorio and Orsingher (2012), lead to the result \eqref{eq:unconddensrefl0}. \end{proof} \begin{remark} For $\frac{d-h}{2}=1,$ that is $d=3$ for $h=1$ and $d=4$ for $h=2$, the function \eqref{eq:unconddensrefl} reduces to \begin{align*} p_h^*(\underline{\bf x}_d,t)&=\frac{\lambda }{c^3t^2\pi^{\frac d2}}\frac{1}{E_{1,\frac d2}(\lambda t)}\left\{ \exp\left\{\lambda t\left(1-\frac{||\underline{\bf x}_{d}||^2}{c^2t^2}\right)\right\}{\bf 1}_{B_R^d}(\underline{{\bf x}}_d)\right.\\ &\quad\left.+\frac{R^{2d}}{||\underline{{\bf x}}_d||^{2d}}\exp\left\{\lambda t\left(1-\frac{R^4}{c^2t^2||\underline{{\bf x}}_d||^2}\right)\right\}{\bf 1}_{C^d_{\frac{R^2}{ct},R}}(\underline{{\bf x}}_d)\right\}, \end{align*} since $E_{1,1}(x)=e^x.$ \end{remark} We observe that the Dirichlet distribution \eqref{eq:dirichlet} with $h=1$ and $d=2$ or $h=2$ and $d=4$, reduces to the uniform law in $S_n$. Therefore, alternatively to \eqref{lawgenpoi}, we can assume that $$g(\underline{\tau}_n;t)=\frac{n!}{t^n}{\bf1}_{S_n}(\underline{\tau}_n).$$ In other words, instead of $\mathcal{N}_d(t)$, we can suppose that the changes of direction are governed by a homogeneous Poisson process $\{N(t),t>0\}$ with rate $\lambda>0$. Under these assumptions, the unconditional distributions of random flights $\{\underline{\bf X}_d(t),t>0\},d=2,4,$ are given by \begin{align}\label{eq:dgo1} p_h(\underline{\bf x}_d,t)&=\sum_{n=1}^\infty p_{n,h}(\underline{\bf x}_d,t)P\{N(t)=n\}\notag\\ &= \begin{cases} \frac{\lambda e^{-\lambda t}}{2\pi c}\frac{e^{\frac{\lambda}{c}\sqrt{c^2t^2-||\underline{\bf x}_2||^2}}}{\sqrt{c^2t^2-||\underline{\bf x}_2||^2}}{\bf 1}_{B_{ct}^2}(\underline{{\bf x}}_2),& d=2,h=1,\\ \frac{\lambda}{c^4t^3\pi^2}e^{-\frac{\lambda}{c^2t}||\underline{\bf x}_4||^2}\left\{2+\frac{\lambda}{c^2t}(c^2t^2-||\underline{\bf x}_4||^2)\right\}{\bf 1}_{B_{ct}^4}(\underline{{\bf x}}_4),& d=4,h=2. \end{cases} \end{align} (for the case $d=2, h=1,$ see (1.2) of Stadje, 1987, (18) of Masoliver {\it et al}., 1993, (20) of Kolesnik and Orsingher, 2005, and for the case $d=4,h=2,$ see formula (3.2) of Orsingher and De Gregorio, 2007). Therefore, if we assume that the changes of direction are governed by the homogeneous Poisson process $\{N(t),t>0\}$, the related reflecting random flights $\{\underline{\bf X}_d^*(t),t>0\},d=2,4,$ have unconditional density functions (in a generalized sense) equal to \begin{align} f_{h}^*(\underline{{\bf x}}_d,t)&=p_h(\underline{{\bf x}}_d,t){\bf 1}_{B_R^d}(\underline{{\bf x}}_d)+\frac{R^{2d}}{||\underline{{\bf x}}_d||^{2d}}p_h\left(R^2\frac{\underline{{\bf x}}_d}{||\underline{{\bf x}}_d||^2},t\right) {\bf 1}_{C^d_{\frac{R^2}{ct},R}}(\underline{{\bf x}}_d)\notag\\ &\quad+\frac{e^{-\lambda t}}{\text{area}\left(\mathbb{S}_{R^2/ct}^{d-1}\right)}\delta_{\{R^2/ct\}}(||\underline{{\bf x}}_d||), \end{align} where $p_h(\underline{{\bf x}}_d,t)$ is given by \eqref{eq:dgo1} and the $\frac{e^{-\lambda t}}{\text{area}\left(\mathbb{S}_{R^2/ct}^{d-1}\right)}\delta_{\{R^2/ct\}}(||\underline{{\bf x}}_d||)$ emerges if $N(t)=0$. \begin{theorem} For $t>\frac Rc$, $0<r\leq R$ and $d=2, h=1,$ we obtain that \begin{align} P\{D_2^*(t)<r\}&=\left[1-\exp\left\{-\lambda t+\frac{\lambda}{c}\sqrt{c^2t^2-r^2}\right\}\right]{\bf 1}_{(0,R]}(r)\notag\\ &\quad+\exp\left\{-\lambda t+\frac{\lambda}{c}\sqrt{c^2t^2-R^4/r^2}\right\}{\bf 1}_{(R^2/ct,R]}(r) \end{align} \end{theorem} \begin{proof} Let us fix $r\in(0,R]$ and $d=2, h=1$. We have that \begin{align*} P\{D_2^*(t)<r\}&= P\{\underline{{\bf X}}_2^*(t)\in B_r^2\}\\ &=\int_{ B_r^2}p_1(\underline{\bf x}_2,t)\de x_1\de x_2+\int_{ C^2_{\frac{R^2}{ct},R}\cap B_r^2}\frac{R^{4}}{||\underline{{\bf x}}_2||^{4}}p_1\left(\frac{R^2}{||\underline{{\bf x}}_2||},t\right)\de x_1 \de x_2. \end{align*} where $p_1\left(\underline{\bf x}_2,t\right)$ is given by \eqref{eq:dgo1}. If $r<R^2/ct$, it is clear that $C^2_{\frac{R^2}{ct},R}\cap B_r^2=\varnothing$ and then \begin{align*} P\{\underline{{\bf X}}_2^*(t)\in B_r^2\} &=\int_{ B_r^2}\frac{\lambda e^{-\lambda t}}{2\pi c}\frac{e^{\frac{\lambda}{c}\sqrt{c^2t^2-||\underline{\bf x}_2||^2}}}{\sqrt{c^2t^2-||\underline{\bf x}_2||^2}}\de x_1\de x_2\\ &=\int_0^r\frac{\lambda e^{-\lambda t}}{c}\rho\frac{e^{\frac{\lambda}{c}\sqrt{c^2t^2-\rho^2}}}{\sqrt{c^2t^2-\rho^2}}\de \rho\\ &=1-\exp\left\{-\lambda t+\frac{\lambda}{c}\sqrt{c^2t^2-r^2}\right\} \end{align*} For $r\in(R^2/ct,R]$, we have that $C^2_{\frac{R^2}{ct},R}\cap B_r^2=C^2_{\frac{R^2}{ct},r}$. Therefore \begin{align*} \int_{ C^2_{\frac{R^2}{ct},r}}\frac{\lambda e^{-\lambda t}}{2\pi c}\frac{R^{4}}{||\underline{{\bf x}}_2||^4}\frac{e^{\frac{\lambda}{c}\sqrt{c^2t^2-R^4/||\underline{{\bf x}}_2||^2}}}{\sqrt{c^2t^2-R^4/||\underline{{\bf x}}_2||^2}}\de x_1 \de x_2&=\int_{R^2/ct}^r\frac{\lambda e^{-\lambda t}}{c}\frac{R^{4}}{\rho^3}\frac{e^{\frac{\lambda}{c}\sqrt{c^2t^2-R^4/\rho^2}}}{\sqrt{c^2t^2-R^4/\rho^2}}\de \rho\\ &=\exp\left\{-\lambda t+\frac{\lambda}{c}\sqrt{c^2t^2-R^4/r^2}\right\}-\exp\{-\lambda t\}. \end{align*} If $R^2/ct<r\leq R$, we also have to consider the discrete part of the distribution of $\{\underline{{\bf X}}_2^*(t),t>0\}.$ Therefore $$P\{D_2^*(t)=R^2/ct\}=P\{\underline{{\bf X}}_2^*(t)\in \mathbb{S}_{R^2/ct}^{d-1}\}=P\{N(t)=0\}=e^{-\lambda t}.$$ This last fact concludes the proof of the theorem. \end{proof} \section{Reflecting random flights on hyperplanes} \subsection{Definitions and probability distributions} In this section we introduce a random flight bouncing off a hyperplane. Let $H(\underline a_d,b):=\{\underline{{\bf x}}_d\in\mathbb{R} ^d:\, \langle\underline a_d,\underline{{\bf x}}_d\rangle=b;\,\underline a_d\in\mathbb{R}^d,b\in\mathbb{R}\}$ be a hyperplane in $\mathbb{R}^d$. A random flight starting from the origin of $\mathbb{R}^d,$ for sufficiently large values of $t$ can be located beyond the hyperplane $H(\underline a_d,b)$. The spherical set of the possible positions $B_{ct}^d$ is therefore composed by the set $L_{ct}^d:=L_{ct}^d(\underline a_d, b):=\{\underline{{\bf x}}_d\in \mathbb{R}^d:||\underline{{\bf x}}_d||^2< c^2t^2,\langle\underline a_d,\underline{{\bf x}}_d\rangle<b\}$ pertaining to the sample paths which have not crossed $H(\underline a_d,b)$ and the set $U_{ct}^d:=U_{ct}^d(\underline a_d, b):=\{\underline{{\bf x}}_d\in \mathbb{R}^d:||\underline{{\bf x}}_d||^2< c^2t^2,\langle\underline a_d,\underline{{\bf x}}_d\rangle\geq b\}$ related to the trajectories which have gone beyond the hyperplane. Of course, if no deviation is recorded by the random flight up to time $t$, the moving particle attains the sphere $\mathbb{S}_{ct}^{d-1},$ which therefore can be split as $\mathbb{S}_{ct}^{d-1}=\partial L_{ct}^d\cup \partial U_{ct}^d,$ where $\partial L_{ct}^d:=\partial L_{ct}^d(\underline a_d, b):=\{\underline{{\bf x}}_d\in \mathbb{R}^d:||\underline{{\bf x}}_d||^2= c^2t^2,\langle\underline a_d,\underline{{\bf x}}_d\rangle<b\}$ and $\partial U_{ct}^d:=\partial U_{ct}^d(\underline a_d, b):=\{\underline{{\bf x}}_d\in \mathbb{R}^d:||\underline{{\bf x}}_d||^2=c^2t^2,\langle\underline a_d,\underline{{\bf x}}_d\rangle\geq b\}$. The reflection of the sample paths crossing $H(\underline a_d,b)$ is described in detail in Appendix \ref{refhyper}. Substantially, the incoming and reflected sample paths form the same angle $\theta$ w.r.t. the normal to the hyperplane. Let $\nu:\mathbb{R}^d\to\mathbb{R}$ be the reflecting (bijective) operator with respect to the hyperplane $H(\underline a_d,b)$ defined as \begin{equation*} \nu(\underline{{\bf x}}_d):=\underline{{\bf x}}_d+2\frac{ b-\langle\underline a_d,\underline{{\bf x}}_d\rangle}{\langle\underline a_d,\underline a_d\rangle}\underline a_d. \end{equation*} Now, we are able to define the reflecting random flight. Let $t':=\inf(t:H(\underline a_d,b)\cap B_{ct}^d\neq \varnothing )$. \begin{definition}\label{defrefhyper} The reflecting random flight $\{\underline{{\bf X}}_d'(t),t>0\},$ reflected by the hyperplane $H(\underline a_d,b)$ is constructed by means of the free process $\{\underline{{\bf X}}_d(t),t>0\}$ as follows: 1) if $t<t'$, then $\underline{{\bf X}}_d'(t)=\underline{{\bf X}}_d(t);$ 2) if $t\geq t'$ and at least one change of direction happens during the time interval $[0,t]$, we have that \begin{align} \underline{{\bf X}}_d'(t)&=\underline{{\bf X}}_d(t){\bf 1}_{L_{ct}^d }(\underline{{\bf X}}_d(t))+\nu(\underline{{\bf X}}_d(t)){\bf 1}_{U_{ct}^d}(\underline{{\bf X}}_d(t)), \end{align} while if no deviation up to time $t$ is recorded \begin{align} \underline{{\bf X}}_d'(t)&=\underline{{\bf X}}_d(t){\bf 1}_{\partial L_{ct}^d }(\underline{{\bf X}}_d(t))+\nu(\underline{{\bf X}}_d(t)){\bf 1}_{\partial U_{ct}^d}(\underline{{\bf X}}_d(t)). \end{align} \end{definition} \begin{figure}[t] \begin{center} \includegraphics[angle=0,width=0.6\textwidth]{reflpaths2.eps} \caption{Four typical sample paths are depicted. The trajectories ${\bf c}$ and ${\bf d}$ are reflected in the set $V_{ct}^d$, while ${\bf e}$ and ${\bf f}$ never cross the reflecting surface $y=b$.}\label{fig2} \end{center} \end{figure} The set of points of $B_{ct}^d$ obtained by reflection $\nu$ around $H(\underline a_d,b)$ is denoted by $V_{ct}^d$, that is $V_{ct}^d:=V_{ct}^d(\underline a_d, b):=\{\underline{{\bf x}}_d\in \mathbb{R}^d:||\nu(\underline{{\bf x}}_d)||^2< c^2t^2,\langle\underline a_d,\underline{{\bf x}}_d\rangle\leq b\}$, while $\partial V_{ct}^d$ stands for the set obtained by the reflection of $\partial U_{ct}^d$ (see Figure \ref{fig2}). The reflection at the hyperplane preserves the form of the sample paths of the free random flight; i.e. the sample paths of $\{\underline{{\bf X}}_d'(t),t>0\}$ are straight lines as well as the trajectories of the free random flights (see Figure \ref{fig2}). Furthermore, the property Q4) guarantees that the reflected trajectories are symmetrically specular w.r.t. the hyperplane $H(\underline a_d,b)$. The conditional distributions of the reflecting random flight $\{\underline{{\bf X}}_d'(t),t>0\}$ are given in the next theorem. \begin{theorem} If $\mathcal N(t)=n$, with $n\geq 1,$ the process $\{\underline{{\bf X}}_d'(t),t>0\},$ has the following conditional density functions \begin{align}\label{eq:condens2} p_{n}'(\underline{{\bf x}}_d,t)=\begin{cases} p_n(\underline{{\bf x}}_d,t){\bf 1}_{B_{ct}^d}(\underline{{\bf x}}_d),& t\leq t'\\ p_n(\underline{{\bf x}}_d,t){\bf 1}_{L_{ct}^d}(\underline{{\bf x}}_d)+p_n\left(\nu(\underline{{\bf x}}_d),t\right){\bf 1}_{V_{ct}^d}(\underline{{\bf x}}_d),& t>t'. \end{cases} \end{align} where $p_n(\underline{{\bf x}}_d,t)$ is equal to \eqref{eq:densfree}. \end{theorem} \begin{proof} The case $t\leq t'$ is trivial. We assume that $t>t'.$ Let $A$ be a Borel set such that $A\cap H(\underline a_d,b)\neq \varnothing$. We observe that \begin{align}\label{eq:step1dens2} P_n\{\underline{{\bf X}}_d'(t)\in A\}&=P_n\{\underline{{\bf X}}_d(t)\in A\cap L_{ct}^d\}+P_n\{\nu(\underline{{\bf X}}_d(t))\in A\cap V_{ct}^d\}\notag\\ &=\int_{ A\cap L_{ct}^d}p_n(\underline{{\bf x}}_d,t)\prod_{k=1}^d\de x_k+P_n\{\nu(\underline{{\bf X}}_d(t))\in A\cap V_{ct}^d\}. \end{align} Let now $$B_A:=\left\{\underline{{\bf y}}_d\in \mathbb{R}^d: \underline{{\bf x}}_d=\nu(\underline{{\bf y}}_d)\in A\cap V_{ct}^d\right\}$$ and thus, by means of Jacobi's transformation formula, we have that \begin{align}\label{eq:step2dens2} P_n\{\nu(\underline{{\bf X}}_d(t))\in A\cap V_{ct}^d\}&=P_n\{\underline{{\bf X}}_d(t)\in B_A\}\notag\\ &=\int_{B_A}p_n(\underline{{\bf y}}_d,t)\prod_{k=1}^d\de y_k\notag\\ &=\int_{A\cap V_{ct}^d} p_n\left(\nu^{-1}(\underline{{\bf x}}_d),t\right)|\text{det}(J_{\nu^{-1}}(\underline{{\bf x}}_d))|\prod_{k=1}^d\de x_k\notag\\ &=\int_{A\cap V_{ct}^d} p_n\left(\nu(\underline{{\bf x}}_d),t\right)\prod_{k=1}^d\de x_k \end{align} where in the last step we have exploited the facts: $\nu=\nu^{-1}$ (which follows from the property Q3)) and $|\text{det}(J_{\nu}(\underline{{\bf x}}_d))|=1$. From \eqref{eq:step1dens2} and \eqref{eq:step2dens2} the result \eqref{eq:condens2} immediately follows. \end{proof} \begin{remark} In view of the property \eqref{eq:normrefl2}, the reflecting random flights introduced by Definition \ref{defrefhyper} are no longer isotropic. Indeed, for $t>t'$, in the density function \eqref{eq:condens2} appears $$p_{n}(\nu(\underline{{\bf x}}_d),t)=p_{n}(||\nu(\underline{{\bf x}}_d)||,t)$$ which does not depend only on the Euclidean distance $||\underline{{\bf x}}_d||.$ \end{remark} Now, we consider reflecting Dirichlet random flights. The assumption \eqref{eq:dirichlet} implies that the reflecting process $\{\underline{{\bf X}}_d'(t),t>0\}$ has probability law \eqref{eq:condens2} given by \begin{align}\label{eq:densrefl2} p_{n,h}'(\underline{{\bf x}}_d,t)&=\frac{\Gamma(\frac{n+1}{2}(d-h)+\frac h2)}{\Gamma(\frac n2(d-h))}\frac{1}{\pi^{\frac d2}(ct)^{(n+1)(d-h)+h-2}}\\ &\quad\times\left[(c^2t^2-||\underline{{\bf x}}_d||^2)^{\frac n2(d-h)-1}{\bf 1}_{L_{ct}^d}(\underline{{\bf x}}_d)+\left(c^2t^2-||\nu(\underline{{\bf x}}_d)||^2\right)^{\frac n2(d-h)-1}{\bf 1}_{V_{ct}^d}(\underline{{\bf x}}_d)\right],\notag \end{align} where $t>t'$ and $||\nu(\underline{{\bf x}}_d)||^2$ is given by \eqref{eq:normrefl2}. In the special cases (i)-(iv) mentioned in Remark \ref{remarkun}, the function \eqref{eq:densrefl2} becomes \begin{align} p_{n,h}'(\underline{{\bf x}}_d,t)&=\frac{1}{\text{area}( \mathbb{S}_{ct}^{d-1})}\left[{\bf 1}_{L_{ct}^d}(\underline{{\bf x}}_d)+{\bf 1}_{V_{ct}^d}(\underline{{\bf x}}_d)\right],\quad t>t'. \end{align} By assuming that the random number of changes of direction has probability law \eqref{lawgenpoi}, for $t>t',$ we have that the unconditional probability distribution of $\{\underline{\bf X}_{d}'(t),t>0\}$ becomes \begin{equation} P\{\underline{\bf X}_{d}'(t)\in \de\underline{\bf x}_{d}\}=p_h'(\underline{\bf x}_d,t)\prod_{k=1}^d\de x_k+\frac{1}{E_{\frac{d-h}{2},\frac d2}(\lambda t)\Gamma(\frac d2)}\mu'(\de\underline{\bf x}_d ) \end{equation} where \begin{align}\label{eq:unconddensrefl} p_h'(\underline{\bf x}_d,t)&=\frac{1}{(ct)^d\pi^{\frac d2}}\frac{1}{E_{\frac{d-h}{2},\frac d2}(\lambda t)}\left\{\left[\gamma_h(||\underline{\bf x}_d||,t)\right]^{1-\frac{2}{d-h}}E_{\frac {d-h}{2},\frac {d-h}{2}}\left(\gamma_h(||\underline{\bf x}_d||,t)\right){\bf 1}_{L_{ct}^d}(\underline{{\bf x}}_d)\right.\notag\\ &\quad+\left.\left[\gamma_h\left(||\nu(\underline{\bf x}_d)||,t\right)\right]^{1-\frac{2}{d-h}}E_{\frac {d-h}{2},\frac {d-h}{2}}\left(\gamma_h\left(||\nu(\underline{\bf x}_d)||,t\right)\right){\bf 1}_{V_{ct}^d}(\underline{{\bf x}}_d)\right\}, \end{align} with $$\gamma_h(||\underline{\bf x}_d||,t):=\lambda t\left(1-\frac{||\underline{\bf x}_{d}||^2}{c^2t^2}\right)^{\frac {d-h}{2}},$$ and $\mu'$ represents the uniform law on $\partial L_{ct}^d\cup \partial V_{ct}^d.$ \begin{remark} It is not hard to prove that the function $$\hat f_\beta(\underline{\bf x}_d,t):=\left(c^2t^2-||\nu(\underline{{\bf x}}_d)||^2\right)^\beta,\quad \underline{\bf x}_d\in V_{ct}^d,\beta\in\mathbb{R},$$ is a solution of the partial differential equation \eqref{eq:epdcon}. Therefore, the second component appearing in \eqref{eq:unconddensrefl}, that is $$\hat p_{n,h}(\underline{{\bf x}}_d,t):=\frac{\Gamma(\frac{n+1}{2}(d-h)+\frac h2)}{\Gamma(\frac n2(d-h))}\frac{\left(c^2t^2-||\nu(\underline{{\bf x}}_d)||^2\right)^{\frac n2(d-h)-1}}{\pi^{\frac d2}(ct)^{(n+1)(d-h)+h-2}},\quad \quad \underline{\bf x}_d\in V_{ct},$$ is a solution of the EPD equation \eqref{eq:epdcon2}. \end{remark} \subsection{On the probability law of the distance from the origin} For the sake of simplicity, we set $\underline a_d=e_d:=(0,...,0,1)$. Therefore, the hyperplane becomes $$H(e_d,b)=\{\underline{{\bf x}}_d\in\mathbb{R} ^d:\, x_d=b;\,b>0\}$$ and the reflection map becomes $$\nu(\underline{{\bf x}}_d):=\underline{{\bf x}}_d+2 (b-x_d),$$with $$||\nu(\underline{{\bf x}}_d)||^2=||\underline{{\bf x}}_d||^2+4b^2-4bx_d. $$ Under the above assumption, we have that \begin{align*} &L_{ct}^d=\{\underline{{\bf x}}_d\in \mathbb{R}^d:||\underline{{\bf x}}_d||^2< c^2t^2,x_d<b\},\\ &V_{ct}^d=\{\underline{{\bf x}}_d\in \mathbb{R}^d:||\underline{{\bf x}}_d||^2+4b^2-4bx_d< c^2t^2,x_d\leq b\}. \end{align*} Let $\{D_d'(t),t>0\}$ where $D_d'(t)=||\underline{{\bf X}}_d'(t)||$. Since the process $\{\underline{{\bf X}}_d'(t),t>0\}$ is not isotropic, the probability distribution of $\{D_d'(t),t>0\}$ is more complicated than \eqref{eq:distancecircinv}. Now, we consider the distribution function $$P_n\{D_d'(t)<r\}$$ with $0<r<ct.$ For $t>t',$ we distinguish the following three cases (see Figure \ref{fig3}): \begin{itemize} \item[1.] $0<r<ct-2b$, where the ball $B_r^d$ does not intersect $V_{ct}^d$; \item[2.] $ct-2b<r<b$, where $B_r^d$ intersects $V_{ct}^d$ but does not overlap $H(e_d,b)$; \item[3.] $b<r<ct$, where $B_r^d$ intersects $V_{ct}^d$ and $H(e_d,b)$. \end{itemize} \begin{figure}[t] \begin{center} \includegraphics[angle=0,width=0.6\textwidth]{reflpathsdistance.eps} \caption{The picture represents cases 1., 2. and 3. emerging in the analysis of $P_n\{D_d'(t)<r\}.$}\label{fig3} \end{center} \end{figure} In the case (i), we simply have that \begin{align} P_n\{D_d'(t)<r\}=P_n\{D_d(t)<r\}=P_n\{\underline{{\bf X}}_d(t)\in B_{r}^d\}. \end{align} In the second case, we must take into account that in $V_{ct}^d\cap B_{r}^d$ we meet reflected sample paths and thus \begin{align} P_n\{D_d'(t)<r\}=P_n\{\underline{{\bf X}}_d(t)\in B_{r}^d\}+P_n\{\nu(\underline{{\bf X}}_d(t))\in V_{ct}^d\cap B_{r}^d\}. \end{align} In the third case $B_r^d=L_r^d\cup U_r^d$ and thus \begin{align}\label{eq:distancethird} P_n\{D_d'(t)<r\}=P_n\{\underline{{\bf X}}_d(t)\in L_{r}\}+P_n\{\nu(\underline{{\bf X}}_d(t))\in V_{ct}^d\cap L_{r}\}. \end{align} The reader should consider that sample paths crossing $x_d=b$ and outside $U_r^d$ can contribute to probability \eqref{eq:distancethird} because the reflected trajectories lie within $V_{ct}^d\cap L_{r}$. All these considerations can be summarized as follows \begin{align} P_n\{D_d'(t)<r\}= \begin{cases} \int_{B_{r}^d}p_{n}(\underline{{\bf x}}_d,t)\prod_{k=1}^d\de x_k,& 0<r<ct-2b,\\ \int_{B_{r}^d}p_{n}(\underline{{\bf x}}_d,t)\prod_{k=1}^d\de x_k+\int_{ V_{ct}^d\cap B_{r}^d}p_{n}(\nu(\underline{{\bf x}}_d),t)\prod_{k=1}^d\de x_k,& ct-2b<r<b,\\ \int_{L_{r}^d}p_{n}(\underline{{\bf x}}_d,t)\prod_{k=1}^d\de x_k+\int_{ V_{ct}^d\cap L_{r}^d}p_{n}(\nu(\underline{{\bf x}}_d),t)\prod_{k=1}^d\de x_k,& b<r<ct. \end{cases} \end{align}
\section{Introduction} Optics-based quantum computing is an attractive possibility. Single-photon source and detector technologies are rapidly maturing \cite{0034-4885-75-12-126503, doi:10.1021/nl404400d, Calkins:13, 1408.1124}, enabling robust photonic-qubit creation and detection. Moreover, single-qubit gates are easily realized with linear optics, and photons are the inevitable carriers for the long-distance entanglement distribution needed to network quantum computers. However, optics-based quantum computing is not without its Achilles' heel, namely the extremely challenging task of realizing a high-fidelity, deterministic, two-qubit entangling gate, such as the {\sc cphase} gate. Knill, Laflamme, and Milburn~\cite{Knill2001} proposed a solution to the preceding two-qubit gate problem by exploiting the nonlinearity afforded by photodetection in conjunction with the introduction of ancilla photons. Their scheme is intrinsically probabilistic, so it requires high-efficiency adaptive measurement techniques and large quantities of ancilla photons to realize useful levels of quantum computation. Consequently, it remains prudent to continue research on more traditional approaches to all-optical two-qubit gates. A prime example is the nonlinear-optical approach first suggested by Chuang and Yamamoto~\cite{PhysRevA.52.3489}, who proposed using Kerr-effect cross-phase modulation (XPM) to impart a $\ensuremath{\text{\foreignlanguage{greek}{p}}}\xspace$-radian phase shift on a single-photon pulse, conditioned on the presence of another single-photon pulse. The fact that the Chuang--Yamamoto architecture provides a deterministic all-optical universal gate set for quantum computation continues to spur work on highly-nonlinear optical fibers \cite{edamatsu, gaeta}, but the single-photon level has yet to be reached. Chuang and Yamamoto's analysis treated the control and target as single-spatiotemporal-mode fields. Later work \cite{PhysRevA.73.062305}, however, examined their architecture using continuous-time XPM theory. Dismissing the possibility of an instantaneous XPM response---owing to its failure to reproduce experimentally-observed classical results---it showed that a causal, non-instantaneous response function introduces fidelity-degrading phase noise which precludes constructing a high-fidelity {\sc cphase} gate. That analysis assumed control and target pulses propagating at the same group velocity, which implied that a uniform single-photon phase shift could not be realized in the fast-response regime, wherein those pulses have durations much longer than that of the XPM response function. Yet fast-response XPM \emph{is} used for imparting uniform conditional phase shifts in fiber-optical switching with classical control pulses \cite{974167}. Those switches' pulses have different group velocities, so that one propagates through the other within the XPM medium. In this paper we develop a continuous-time quantum XPM theory for pulses with differing group velocities \cite{footnote3}, and then use it to assess the feasibility of extending the fiber-switching technique to the single-photon regime for creating a {\sc cphase} gate. We show that causality-induced phase noise still rules out high-fidelity $\ensuremath{\text{\foreignlanguage{greek}{p}}}\xspace$-radian conditional phase shifts for both a reasonable theoretical model for the XPM response function and for the experimentally-measured XPM response function of silica-core fiber. \section{Quantum XPM Theory} Our theory begins with classical XPM for a pair of single-spatial-mode continuous-time scalar fields---with center frequencies $\omega_A$ and $\omega_B$ and complex envelopes $E_A(z,t)$ and $E_B(z,t)$---that propagate from $z=0$ to $z=L$ through an XPM medium. Because we are interested in ultimate limits on the utility of XPM for two-qubit gates, we neglect loss, dispersion, and self-phase modulation. Thus the behavior of the classical complex envelopes of interest is governed by the coupled-mode equations \cite{agrawal2001nonlinear}: \begin{subequations} \label{eq:coupledmode} \begin{align} \left( \frac{\partial}{\partial{z}} + \frac{1}{v_A} \frac{\partial}{\partial{t}} \right) E_{A}(z,t) &= \text{i} n_A(z,t) E_A(z,t), \\ \left( \frac{\partial}{\partial{z}} + \frac{1}{v_B} \frac{\partial}{\partial{t}} \right) E_{B}(z,t) &= \text{i} n_B(z,t) E_B(z,t). \end{align} \end{subequations} Here, $v_A$ and $v_B$, satisfying $v_B > v_A$, are the group velocities of $E_A(z,t)$ and $E_B(z,t)$, and $n_A(z,t)$ and $n_B(z,t)$ are the intensity-dependent refractive indices that these fields encounter. For convenient linking to the quantum analysis, we normalize $E_A(z,t)$ and $E_B(z,t)$ to make $\hbar\omega_{K}I_{K}(z,t) =\hbar\omega_{K}|E_{K}(z,t)|^2,$ for $K= A,B$, the powers carried by these fields. The nonlinear refractive indices are then given by \cite{footnote4} \begin{subequations} \begin{align} n_A(z,t) &= \eta \int_{-\infty}^t \dd{t'} h(t-t') I_B(z,t'), \\ n_B(z,t) &= \eta \int_{-\infty}^t \dd{t'} h(t-t') I_A(z,t'), \end{align} \end{subequations} where $\eta$ is the strength of the nonlinearity and $h(t)$ is its real-valued, causal response function, normalized to satisfy $\int_0^\infty \dd{t} h(t) = 1$. In the quantum theory for the preceding XPM setup, $E_A(z,t)$ and $E_B(z,t)$ become baseband field operators, $\hat{E}_A(z,t)$ and $\hat{E}_B(z,t)$ with units $\sqrt{\rm photons/s}$. At the input and output planes, $z=0$ and $z=L$, these field operators must satisfy the canonical commutation relations for free fields, viz., \begin{subequations} \label{commutators} \begin{align} [\hat{E}_K(z,t),\hat{E}_J(z,s)] &= 0,\\ [\hat{E}_K(z,t),\hat{E}_J^\dagger(z,s)] &= \delta_{JK}\delta(t-s), \end{align} \end{subequations} for $K = A,B$, $J = A,B$, and $z=0,L$. Unless $h(t)=\delta(t)$, which \cite{PhysRevA.73.062305} has ruled out for its failure to reproduce experimentally-observed classical results, Langevin noise terms must be added to the classical coupled-mode equations to ensure that the output fields have the required commutators. Here, we take a cue from the work of Boivin \em et al\/\rm.~\cite{PhysRevLett.73.240}, which developed a continuous-time quantum theory of self-phase modulation and which \cite{PhysRevA.73.062305} extended to XPM when both fields have the same group velocity. The quantum coupled-mode equations that result are \begin{subequations} \label{eq:quantumcoupledmode} \begin{align} \left( \frac{\partial}{\partial{z}} + \frac{1}{v_A} \frac{\partial}{\partial{t}} \right) \hat{E}_{A}(z,t) &= i [\hat{n}_A(z,t) + \hat{m}_A(z,t)] \hat{E}_A(z,t), \\ \left( \frac{\partial}{\partial{z}} + \frac{1}{v_B} \frac{\partial}{\partial{t}} \right) \hat{E}_{B}(z,t) &= i [\hat{n}_B(z,t) + \hat{m}_A(z,t)] \hat{E}_B(z,t). \end{align} \end{subequations} In terms of the photon-flux operators $\hat{I}_{K}(z,t)\equiv \hat{E}_{K}^\dagger(z,t)\hat{E}_{K}(z,t)$, for $K= A,B$, the nonlinear refractive indices are now operator-valued and given by \begin{subequations} \begin{align} \hat{n}_A(z,t) &= \eta \int_{-\infty}^t \dd{t'} h(t-t') \hat{I}_B(z,t'), \\ \hat{n}_B(z,t) &= \eta \int_{-\infty}^t \dd{t'} h(t-t') \hat{I}_A(z,t'). \end{align} \end{subequations} The Langevin noise operators $\hat{m}_A(z,t)$ and $\hat{m}_B(z,t)$ are \begin{subequations} \begin{eqnarray} \lefteqn{\hat{m}_A(z,t) = \int_{0}^\infty \frac{\dd{\omega}}{2\ensuremath{\text{\foreignlanguage{greek}{p}}}\xspace} \sqrt{\eta H_\text{im}(\omega)} } \nonumber \\ && \times\,\{[\hat{B}(z,\omega) - \text{i} \hat{C}^\dagger(z,\omega)] \text{e}^{-\text{i}\omega t} + \textrm{hc}\}, \\[.05in] \lefteqn{\hat{m}_B(z,t) = \int_{0}^\infty \frac{\dd{\omega}}{2\ensuremath{\text{\foreignlanguage{greek}{p}}}\xspace} \sqrt{\eta H_\text{im}(\omega)} } \nonumber \\ && \times\,\{[\hat{B}(z,\omega) + \text{i} \hat{C}^\dagger(z,\omega)] \text{e}^{-\text{i}\omega t} + \textrm{hc}\}, \end{eqnarray} \end{subequations} where $H_\text{im}(\omega)$ is the imaginary part of the frequency response $H(\omega) = \int_0^\infty \dd{t} h(t) e^{i \omega t}$, $\hat{B}(z,\omega)$ and $\hat{C}(z,\omega)$ are independent frequency-domain bosonic field operators \cite{footnote1} taken to be in thermal states at absolute temperature $T$, and $\textrm{hc}$ denotes the Hermitian conjugate. Equation~(\ref{eq:quantumcoupledmode}) can be solved to yield the following input-output relations \begin{subequations}\label{eq:io} \begin{align} \hE^\text{out}_A(t) &= \text{e}^{\text{i}\hat{\xi}_A(t)} \text{e}^{\text{i}\hat{\zeta}_A(t)} \hE^\text{in}_A(t), \\ \hE^\text{out}_B(t) &= \text{e}^{\text{i}\hat{\xi}_B(t)} \text{e}^{\text{i}\hat{\zeta}_B(t)} \hE^\text{in}_B(t)\text{,} \end{align} \end{subequations} for the output field operators, $\hE^\text{out}_K(t)\equiv\hat{E}_K(L,t+L/v_K)$, in terms of the input field operators, $\hE^\text{in}_K(t)\equiv\hat{E}_K(0,t)$, the \emph{phase-shift operators} \cite{footnote2} \begin{subequations} \begin{align} \hat{\zeta}_A(t) \equiv \eta \int_0^{L} \dd{z} \int \dd{s} h(t-s) \hI^\text{in}_B(s+z/u), \\ \hat{\zeta}_B(t) \equiv \eta \int_{0}^{L} \dd{z} \int \dd{s} h(t-s) \hI^\text{in}_A(s-z/u), \end{align} \end{subequations} where $\hI^\text{in}_K(t) \equiv {\hE^{{\text{in}}\dagger}_K}(t)\hE^\text{in}_K(t)$ and $1/u \equiv 1/v_A - 1/v_B$, and the \emph{phase-noise operators} \begin{subequations} \begin{eqnarray} \lefteqn{\hat{\xi}_A(t) \equiv \int_0^L\dd{z}\int_{0}^{\infty} \frac{\dd{\omega}}{2\pi} \sqrt{\eta H_\text{im}(\omega) }} \nonumber \\ && \times\,\{ [\hat{B}(z,\omega) - i \hat{C}^{\dagger}(z,\omega)]e^{-i \omega (t+z/v_A)} + \textrm{hc}\}, \\[.05in] \lefteqn{\hat{\xi}_B(t) \equiv \int_0^L\dd{z}\int_{0}^{\infty} \frac{\dd{\omega}}{2\pi} \sqrt{\eta H_\text{im}(\omega) }} \nonumber \\ && \times\,\{ [\hat{B}(z,\omega) + i \hat{C}^{\dagger}(z,\omega)]e^{-i \omega (t+z/v_B)} + \textrm{hc}\}. \end{eqnarray} \end{subequations} These input-output relations ensure that $\hE^\text{out}_A(t)$ and $\hE^\text{out}_B(t)$ have the proper free-field commutators, as required by Eq.~(\ref{commutators}). The phase-noise operators have nonzero commutator \begin{eqnarray} \lefteqn{[\hat{\xi}_A(t) , \hat{\xi}_B(s) ] = } \nonumber \\ && \ \text{i} \eta \int_0^{L} \dd{z} [h(s-t-z/u) - h(t-s+z/u)], \end{eqnarray} and they are in a zero-mean jointly Gaussian state that is characterized by the symmetrized autocorrelation functions \begin{equation} \braket{\hat{\xi}_K(t)\hat{\xi}_K(s) + \hat{\xi}_K(s)\hat{\xi}_K(t)} = \int\frac{\dd{\omega}}{\pi} S_{\xi\xi}(\omega)\cos[\omega(t-s)]\text{,} \label{eq:xi3} \end{equation} for $K=A,B$, with spectrum \begin{equation}\label{eq:xi4} S_{\xi\xi}(\omega) = \eta H_\text{im}(\omega) \coth\left(\frac{\hbar\omega}{2k_BT}\right)\text{,} \end{equation} where $k_B$ is the Boltzmann constant. For the theory to make physical sense, it must be that $H_\text{im}(\omega) \geq 0$ for all $\omega\ge 0$~\cite{Shapiro:97, PhysRevLett.73.240}, because noise spectra must be nonnegative. \section{XPM-Based CPHASE Gate} To build a {\sc cphase} gate from the preceding quantum XPM interaction we proceed as follows. Consistent with dual-rail logic \cite{PhysRevA.52.3489}, the input and output field operators are chosen to be in states in the Hilbert space spanned by their computational basis states, $\{|0\rangle_K, |1\rangle_K : K = A,B\}$. We will take $\ket{0}_K$ to be the vacuum state, and set \begin{align} \ket{1}_K = \int \dd{t} \psi_K(t) \ket{t}_K\text{,} \end{align} where the wave functions $\{\psi_K(t) : K = A,B\}$ are normalized ($\int \dd{t} |\psi_K(t)|^2 = 1$), and $\ket{t}_K$ is the state of $\hE^\text{in}_K(t)$ or $\hE^\text{out}_K(t)$ in which there is a single photon at time $t$ and none at all other times. To enforce the interchangeability of the control and target qubits, we take the single-photon pulses in each field to have the same pulse shape. Moreover, because we have assumed $v_B > v_A$, we will assume that $\psi_B(t) = \psi_A(t - t_d)$, where $t_d > 0$ is a delay, specified below, chosen to allow the single-photon excitation in $\hE^\text{in}_B(t)$ to propagate through the one in $\hE^\text{in}_A(t)$ while both are within the nonlinear medium, thus ensuring each imposes a uniform phase shift on the other. Sufficient conditions for guaranteeing a uniform phase shift are intuitive and easily derived. Ignoring the phase noise for now, the phase shifts induced on each field by the presence of a single-photon pulse in the other field are found by taking the partial trace of the phase-shift operator for one field with respect to the other: \begin{subequations} \begin{align} \tensor[_B]{\braket{1 | \text{e}^{\text{i}\hat{\zeta}_A(t)} | 1}}{_B} = & \int \dd{s} \text{e}^{\text{i} \eta\int^L_0\dd{z} h(t-s+z/u)} |\psi_B(s)|^2, \label{eq:induced-phase-shift_a}\\ \tensor[_A]{\braket{1 | \text{e}^{\text{i}\hat{\zeta}_B(t)} | 1}}{_A} = & \int \dd{s} \text{e}^{\text{i} \eta\int^L_0\dd{z} h(t-s-z/u)} |\psi_A(s)|^2. \label{eq:induced-phase-shift_b} \end{align} \end{subequations} From these expressions it is clear that a sufficient condition for a uniform phase shift on $\hE^\text{out}_K(t)$ is that the response-function integrals, i.e., the integrals in the exponents in Eqs.~(\ref{eq:induced-phase-shift_a}) and (\ref{eq:induced-phase-shift_b}), encapsulate the entirety of the response function for all times $t$ and $s$ for which $\psi_K(t)$ and $\psi_J(s)$ are nonzero, where $J \neq K$. Suppose that $h(t)$, $\psi_A(t)$, and $\psi_B(t)$ are only nonzero over the intervals $[0,t_h]$, $[-t_\psi/2, t_\psi/2]$, and $[-t_\psi/2 + t_d, t_\psi/2 + t_d]$, respectively. Although these conditions might not be satisfied exactly, e.g., when the response function and/or the pulse shapes do not have bounded support, we can at least take $t_h$ and $t_\psi$ to represent the nominal durations over which each function is significantly different from zero. In terms of these time durations, our sufficient conditions for uniform phase shifts are \begin{subequations} \begin{align} t_d & \geq t_\psi + t_h \label{eq:uniform-phase-1_a}\\ \frac{L}{u} & \geq t_\psi + t_h + t_d. \label{eq:uniform-phase-1_b} \end{align} \end{subequations} Under these conditions, we have that \begin{align} \int^L_0\dd{z} h(t-s+z/u) = \int^L_0\dd{z} h(t-s-z/u) = u, \end{align} in Eqs.~(\ref{eq:induced-phase-shift_a}) and (\ref{eq:induced-phase-shift_b}), respectively, due to the normalization of $h(t)$. Hence these equations reduce to \begin{align} \tensor[_B]{\braket{1 | \text{e}^{\text{i}\hat{\zeta}_A(t)} | 1}}{_B} = \tensor[_A]{\braket{1 | \text{e}^{\text{i}\hat{\zeta}_B(t)} | 1}}{_A} = \text{e}^{\text{i} \eta u}, \end{align} after making use of the wave functions' normalization. From this result we see that the uniform phase shift imposed by the presence of a single-photon excitation is \begin{align}\label{eq:phase-magnitude} \phi = \eta u, \end{align} so that the strength, $\eta$, of the XPM nonlinearity must be $\eta = \ensuremath{\text{\foreignlanguage{greek}{p}}}\xspace/u$ to realize the desired {\sc cphase} gate Physically, the condition in Eq.~(\ref{eq:uniform-phase-1_a}) implies that the fast pulse, $\hat{E}_B(z,t)$, does not enter the XPM medium until the entirety of the slow pulse, $\hat{E}_A(z,t)$, and its nonlinear response have propagated into it. Similarly, Eq.~(\ref{eq:uniform-phase-1_b}) implies that the slow pulse, $\hat{E}_A(z,t)$, does not exit the XPM medium until the entirety of the fast pulse, $\hat{E}_B(z,t)$, and its nonlinear response have propagated out of it. Taking both the delay and the XPM medium to be as short as possible, which will prove most favorable with regards to the phase noise, these conditions simplify to \begin{subequations}\label{eq:uniform-phase-2} \begin{align} t_d & = t_\psi + t_h \\ \frac{L}{u} & = 2(t_\psi + t_h). \end{align} \end{subequations} \section{Vacuum and Single-Photon Fidelities} A complete fidelity analysis for the preceding XPM-based {\sc cphase} gate would evaluate the overlap between the actual two-field output state from the XPM interaction and the two-field output state from an ideal {\sc cphase} gate, averaged uniformly over all possible two-field input states. We, however, will limit our attention to the vacuum and single-photon fidelities, introduced in \cite{PhysRevA.73.062305}. Let $\hE^\text{in}_B(t)$, regarded as the gate's control field, be in its vacuum (single-photon) state. The ensuing vacuum (single-photon) fidelity $F_0$ ($F_1$) is the overlap between the actual state for $\hE^\text{out}_A(t)$ and that field's ideal state, averaged uniformly over all $\hE^\text{in}_A(t)$ states on the Bloch sphere. The phase-noise limits we will find for these two fidelities are closely related, so let us begin with $\hE^\text{in}_B(t)$ being in its vacuum state, in which case the ideal {\sc cphase}-gate state for $\hE^\text{out}_A(t)$ is its input state. The formula from \cite{PhysRevA.73.062305} for the vacuum fidelity is \begin{eqnarray} \lefteqn{F_0 = \frac{1}{3} \left [1 + {\rm Re}\left( \int \dd{t} \braket{\text{e}^{\text{i}\hat{\xi}_A(t)}} |\psi_A(t)|^2 \right) \right.} \nonumber \\[.05in] &&\hspace*{-.1in}+\,\left. \int \dd{t} \int \dd{s} |\psi_A(t)|^2 |\psi_A(s)|^2 \braket{\text{e}^{\text{i}[\hat{\xi}_A(t) -\hat{\xi}_A(s)]}} \right]. \label{eq:vac-fid} \end{eqnarray} We can place an upper bound on the vacuum fidelity by letting $T=0\,\text{K}$, to minimize the phase noise, and setting the third term in Eq.~(\ref{eq:vac-fid}) to 1, to maximize its value. Using the $\braket{e^{i\hat{\xi}_A(t)}}$ value for the phase noise's zero-mean Gaussian state whose spectrum is given by Eq.~(\ref{eq:xi4}) with $T=0$\,K, we find that \begin{align} F_0 \le \frac{2}{3} + \frac{1}{3} \exp\!\left(-\frac{\eta L}{4\ensuremath{\text{\foreignlanguage{greek}{p}}}\xspace} \int \dd{\omega}|H_\text{im}(\omega)|\right). \label{nonuniformF0bound} \end{align} Under our uniform-phase-shift conditions, Eqs.~(\ref{eq:uniform-phase-2}) and (\ref{eq:phase-magnitude}), this bound on the vacuum fidelity becomes \begin{align} \label{fidelitybound} F_0 \le \frac{2}{3} + \frac{1}{3} \exp\!\left(-\frac{\phi}{2\ensuremath{\text{\foreignlanguage{greek}{p}}}\xspace}(t_\psi + t_h) \int \dd{\omega}|H_\text{im}(\omega)|\right). \end{align} It is readily apparent from (\ref{fidelitybound}) that the vacuum fidelity decreases as the phase shift increases. Likewise, it is clear that perfect fidelity for a nonzero phase shift is impossible, \emph{even in theory}, for any physically-valid response function. Perfect fidelity for a nonzero phase-shift $\phi$ requires either $t_\psi+t_h = 0$ or $|H_\text{im}(\omega)| = 0$ for all $\omega$. The former is impossible for non-instantaneous pulse shapes and response functions, while the latter is impossible for non-instantaneous, causal response functions. An even looser, more favorable bound can be gotten by presuming operation to be in the slow-response regime, wherein $t_\psi \ll t_h$. For a $\ensuremath{\text{\foreignlanguage{greek}{p}}}\xspace$-radian phase shift, we are then left with \begin{align} F_0 \le F_\text{max} \equiv \frac{2}{3} + \frac{1}{3} \exp\!\left(-(t_h/2) \int \dd{\omega}|H_\text{im}(\omega)|\right). \label{eq:vac-fid-max} \end{align} This $F_\text{max}$ result also applies to the single-photon fidelity, whose general expression is \cite{PhysRevA.73.062305}, \begin{align} F_1 &= \frac{1}{3} \left [1 + {\rm Re}\left( \text{e}^{-i\phi}\int \dd{t} \braket{\text{e}^{\text{i}\hat{\xi}_A(t)}}\braket{\text{e}^{\text{i}\hat{\zeta}_A(t)}} |\psi_A(t)|^2 \right) \right. \nonumber \\[.05in] &+\,\int \dd{t} \int \dd{s} |\psi_A(t)|^2 |\psi_A(s)|^2 \braket{\text{e}^{\text{i}[\hat{\xi}_A(t) -\hat{\xi}_A(s)]}} \nonumber \\[.05in] &\times \left.\braket{\text{e}^{\text{i}[\hat{\zeta}_A(t)-\hat{\zeta}_A(s)]}} \right], \label{eq:singlephoton-fid} \end{align} which reduces to the result in Eq.~(\ref{eq:vac-fid}) when the XPM interaction produces a uniform $\phi$-radian phase shift. Assuming $\phi = \ensuremath{\text{\foreignlanguage{greek}{p}}}\xspace$ and operation in the slow-response regime, we then get \begin{align} F_1 \le \frac{2}{3} + \frac{1}{3} \exp\!\left(-(t_h/2) \int \dd{\omega}|H_\text{im}(\omega)|\right) \label{eq:singlephoton-fid-max} \end{align} from Eq.~(\ref{eq:singlephoton-fid}), thus putting the same optimistic but likely unobtainable upper limit on both the vacuum and single-photon fidelities for an XPM-based gate that produces uniform $\ensuremath{\text{\foreignlanguage{greek}{p}}}\xspace$-radian phase shifts. \section{Principal Mode Projection} The fidelity upper limit we have found for both the vacuum and single-photon fidelities increases with decreasing phase shift, so a natural question arises: Can we cascade a series of small-phase-shift gates, interspersed with quantum error correction, to realize a high-fidelity {\sc cphase} gate? The errors addressed by quantum-computation error correction---dephasing noise, depolarizing noise, bit flips, etc.---all lie \emph{within} the Hilbert space for the qubits of interest \cite{nielsen2000quantum}. In our case, however, phase noise randomly distorts the single-photon pulse shape while it preserves photon flux, $\hI^\text{in}_K(t)=\hI^\text{out}_K(t)$, so there is no photon loss. Thus it causes the state to drift \emph{outside} the computational Hilbert space, rendering traditional quantum-error-correction techniques of no value. An alternative approach would be to reshape the pulses after each XPM interaction, but the random nature of the phase noise precludes this approach's succeeding. Instead, let us pursue the route of principal-mode projection (PMP), as suggested in~\cite{PhysRevA.87.042325}. There, a $\vee$-type atomic system in a one-sided cavity was part of a unit cell comprising the atomic nonlinearity followed by filtering to project its output onto the computational-basis temporal mode (the \em principal\/\rm\ mode). Cascading a large number of these unit cells---each producing a small phase shift but with an even smaller error---yielded the $\ensuremath{\text{\foreignlanguage{greek}{p}}}\xspace$-radian phase shift needed for a {\sc cphase} gate with a fidelity that, in principle, could be arbitrarily high if enough unit cells were employed. It behooves us to see whether a similar favorable error versus phase-shift tradeoff applies to our XPM system. Sadly, as we now show, such is not the case. Consider a single iteration of XPM+PMP when $\hE^\text{in}_A(t)$ is in state $\alpha\ket{0}_A+\beta\ket{1}_A$, with $|\alpha|^2 + |\beta|^2 = 1$, {and $\hE^\text{in}_B(t)$ is in its vacuum state. The density operator for $\hE^\text{out}_A(t)$ will then be \begin{align} \hat{\rho}^{(0)}_{\text{PMP}} &= (1 - |\beta|^2\braket{|\hat{\mathcal{T}}|^2}) \ket{0}_A\!\bra{0} + \alpha\beta^*\braket{\hat{\mathcal{T}}^\dagger} \ket{0}_A\!\bra{1} \nonumber \\ &\,\,+ \alpha^*\beta\braket{\hat{\mathcal{T}}}\ket{1}_A\!\bra{0} + |\beta|^2 \braket{|\hat{\mathcal{T}}|^2} \ket{1}_A\!\bra{1},\label{eq:pmp-out0} \end{align} where $\hat{\mathcal{T}}\equiv\int \dd{t} |\psi_A(t)|^2 \text{e}^{\text{i} \hat{\xi}_A(t)}$ can be thought of as the photon-flux transmissivity of the abstract pulse-shape filter responsible for carrying out the PMP. If the XPM interaction produces a uniform $\phi$-radian phase shift, then the \em same\/\rm\ expression gives the density operator for $\hE^\text{out}_A(t)$ when $\hE^\text{in}_B(t)$ is in its single-photon state $|1\rangle_B$. Consequently, after averaging $\{\alpha, \beta\}$ over the Bloch sphere we find that the vacuum and single-photon fidelities satisfy \begin{align} F_0 &= F_1 = \frac{1}{2} + \frac{1}{3}\braket{{\rm Re}(\hat{\mathcal{T}})} + \frac{1}{6}\braket{|\hat{\mathcal{T}}|^2}\label{eq:pmp-vac-fid}\\ &= \frac{1}{2} + \frac{1}{3} \braket{\text{e}^{\text{i} \hat{\xi}_A(t)}} \nonumber \\[.05in] &\,\,+ \frac{1}{6}\int \dd{t} \int \dd{s} |\psi_A(t)|^2 |\psi_A(s)|^2 \braket{\text{e}^{\text{i}[\hat{\xi}_A(t)-\hat{\xi}_A(s)]}}, \label{pmp-fid} \end{align} where we have used the fact that $\braket{\text{e}^{\text{i} \hat{\xi}_A(t)}}$ is constant and real-valued. Comparing this result to Eq.~(\ref{eq:vac-fid}), we see that a single iteration of PMP \emph{does} increase both $F_0$ and $F_1$, but it does \em not\/\rm\ increase $F_\text{max}$ from what is given in (\ref{eq:vac-fid-max}), a bound that still applies to both the vacuum and single-photon fidelities. Now it is easy to see that cascading $N$ unit cells of XPM+PMP cannot avoid the fidelity limit identified in the previous section. For such a cascade, $F_0$ and $F_1$ obey Eq.~(\ref{eq:pmp-vac-fid}) with $\hat{\mathcal{T}}$ replaced by $\prod_{n=1}^N\hat{\mathcal{T}}_n$, where $\hat{\mathcal{T}}_n\equiv\int \dd{t} |\psi_A(t)|^2 \text{e}^{\text{i} \hat{\xi}_{A_n}(t)}$ is the photon-flux transmissivity of the $n$th XPM+PMP unit cell. But, the $\{\hat{\xi}_{A_n}(t)\}$ are statistically independent and identically distributed, so that Eq.~(\ref{pmp-fid}) for the $N$ unit-cell cascade is then \begin{align} F_0 &= F_1 = \frac{1}{2} + \frac{1}{3} \prod_{n=1}^N\braket{\text{e}^{\text{i} \hat{\xi}_{A_n}(t)}} \nonumber \\[.05in] &\,\,+ \frac{1}{6}\int \dd{t} \int \dd{s} |\psi_A(t)|^2 |\psi_A(s)|^2 \prod_{n=1}^N\braket{\text{e}^{\text{i}[\hat{\xi}_{A_n}(t)-\hat{\xi}_{A_n}(s)]}} \\[.05in] &\le \frac{2}{3} + \frac{1}{3}\exp\!\left(-(t_h/2) \int \dd{\omega}|H_\text{im}(\omega)|\right), \label{PMPbound} \end{align} where the inequality is obtained by assuming that each XPM+PMP unit cell operates in the slow-response regime and provides a uniform phase shift of $\ensuremath{\text{\foreignlanguage{greek}{p}}}\xspace/N$. That this fidelity bound coincides with $F_\text{max}$ for a single XPM interaction that produces a uniform $\ensuremath{\text{\foreignlanguage{greek}{p}}}\xspace$-radian phase shift is a consequence of quantum XPM's phase shift and error scaling identically with the nonlinearity's strength, $\eta$. \section{Fiber-XPM Fidelity Bounds} In this section we will evaluate the fidelity bound $F_\text{max}$ for two XPM response functions: a reasonable theoretical model and the experimentally-measured response function of silica-core fiber. We start with the family of single-resonance, two-pole response functions characterized by the frequency response \begin{align}\label{eq:response} H(\omega) = \frac{\omega_0^2}{\omega_0^2 - \omega^2 - \text{i} \omega \gamma}. \end{align} This family, which was employed in~\cite{PhysRevA.73.062305}, includes a common approximation to the Raman response function of silica-core fiber \cite{Lin:06}. For $0 < \gamma/2 < \omega_0$, its response function $h(t)$ is underdamped, \begin{align} h(t) = \frac{\omega_0^2\text{e}^{-\gamma t/2} \sin\left(\sqrt{\omega_0^2 - \gamma^2/4}\,t \right)}{\sqrt{\omega_0^2 - \gamma^2/4}} \text{, for } t \geq 0; \end{align} for $\gamma/2=\omega_0$ it is critically damped, \begin{align} h(t) = \omega_0^2 t \text{e}^{-\omega_0t} \text{, for } t \geq 0; \end{align} and for $\gamma/2 > \omega_0$ it is overdamped, \begin{align} h(t) = \frac{\omega_0^2\text{e}^{-\gamma t/2} \sinh\left(\sqrt{\gamma^2/4 - \omega_0^2}\, t \right)}{\sqrt{\gamma^2/4 - \omega_0^2}} \text{, for } t \geq 0. \end{align} In all of these cases $h(t)$ has infinite duration, so we will optimistically take $t_h$ to be the root-mean-square duration of $h(t)$, \begin{align} t_h & = \sqrt{\frac{\int_0^\infty \dd{t} t^2 h^2(t)}{\int_0^\infty \dd{t} h^2(t)} - \left(\frac{\int_0^\infty \dd{t} t h^2(t)}{\int_0^\infty \dd{t} h^2(t)} \right) ^2}\\ & = \sqrt{\frac{1}{\gamma^2} + \frac{\gamma^2}{4\omega_0^4} - \frac{1}{2\omega_0^2}}, \end{align} which satisfies \begin{align} \omega_0 t_h = \sqrt{\frac{1}{\Gamma^2} + \frac{\Gamma^2}{4} - \frac{1}{2}}, \end{align} in terms of the dimensionless parameter $\Gamma = \gamma/\omega_0$. This duration is minimized at $\Gamma = \sqrt{2}$, which is slightly into the underdamped regime. In terms of $\Gamma$, it can be shown that \begin{align}\label{eq:arctan} \frac{\int \dd{\omega}|H_\text{im}(\omega)|}{\omega_0} & = \frac{\ensuremath{\text{\foreignlanguage{greek}{p}}}\xspace \text{i} + 2 \tanh ^{-1}\left(\frac{\Gamma ^2-2}{\Gamma \sqrt{\Gamma ^2-4}}\right)}{\sqrt{\Gamma ^2-4}}, \end{align} which makes it easy to evaluate $F_\text{max}$, as a function of $\Gamma$, from Eq.~(\ref{eq:vac-fid-max}), as shown in Fig.~\ref{fig:fidelity}. Note that, despite its appearance, the expression on the right in Eq.~(\ref{eq:arctan}) is real-valued for $\Gamma\geq 0$. \begin{figure} \centering \includegraphics[scale=0.3]{fidelity-plot.eps} \caption{Fidelity upper-bound $F_\text{max}$ for the single-resonance, two-pole response function plotted versus the normalized damping parameter $\Gamma$.} \label{fig:fidelity} \end{figure} Figure~\ref{fig:fidelity} shows that $F_\text{max}$ peaks at just less than 82\%. It is worth emphasizing, in this regard, that $F_\text{max}$ is a very generous upper bound: (1) it does not include the effects of loss, dispersion, or self-phase modulation; (2) it assumes operation at $T=0$\,K; (3) it assumes operation in the slow-response regime, which would imply $\psi_A(t)$ and $\psi_B(t)$ had sub-femtosecond durations; (4) its use of $h(t)$'s root-mean-square duration for $t_h$ is an optimistic value insofar as uniform phase-shift conditions are concerned; and (5) it has generously set the third term of Eq.~(\ref{eq:vac-fid}) to its upper limit of 1. Accordingly, it seems fair to say that, at least for this response function, fiber XPM will not lead to a high-fidelity {\sc cphase} gate. \begin{figure} \centering \subfloat[Frequency response.]{ \includegraphics[scale=0.35]{fiber-freq-response.eps}}\\ \subfloat[Temporal response.]{ \includegraphics[scale=0.35]{fiber-time-response.eps}} \caption{\label{fig:fiber-response} The Raman response of silica-core fiber, as measured by Stolen~et~al.~\cite{Stolen:89}.} \end{figure} At this point we could continue by evaluating the behavior of $F_\text{max}$ for other idealized theoretical response functions, but it is better to employ the XPM response function of fused-silica fiber. XPM-based fiber-optical switches typically employ co-polarized inputs \cite{974167}, and for such inputs that response function is the fiber's co-polarized Raman response function \cite{Lin:06}, which was measured by Stolen~\em et~al\/\rm.~\cite{Stolen:89} and is shown in Fig.~\ref{fig:fiber-response}. For this response we have that the root-mean-square duration of $h(t)$ is $t_h\approx 49.2\,\text{fs}$ and \begin{align} \int \dd{\omega}|H_\text{im}(\omega)| \approx 1.79\times 10^{14}\,\text{rad}/\text{s}. \end{align} These values imply that $F_\text{max}\approx 67.1\%$, which is worse than what we found for the $\Gamma$-optimized two-pole response. If we stick with the Raman response function, we can explore $F_1$ fidelity behavior when we relax our uniform phase-shift condition. In particular, our uniform-phase-shift conditions make good sense when the pulse width is significant relative to the response function's duration. However, deep in the slow-response regime---which these very same conditions suggest is optimal---the $\hE^\text{in}_A(t)$ and $\hE^\text{in}_B(t)$ pulse shapes are well approximated by Dirac-delta distributions relative to the response function. Thus it would seem that ensuring the entirety of the pulse be exposed to the entirety of the response is not particularly critical, in this regime, as there is very little pulse to begin with. Suppose we are aiming for a $\ensuremath{\text{\foreignlanguage{greek}{p}}}\xspace$-radian phase shift. Then, presuming operation at $T=0$\,K \em without\/\rm\ imposing the uniform phase-shift conditions, the vacuum and single-photon fidelities are bounded by (\ref{nonuniformF0bound}) for $F_0$, and \begin{align} F_1 &\le F_1^\text{max} \equiv\frac{2}{3} - \frac{1}{3} \exp\!\left(-\frac{\eta L}{4\ensuremath{\text{\foreignlanguage{greek}{p}}}\xspace}\int \dd{\omega} H_\text{im}(\omega)\right)\nonumber \\[.05in] &\times {\rm Re}\left( \int \dd{t} \int \dd{s} \text{e}^{\text{i} \eta\int^L_0\dd{z} h(t-s+z/u)}|\psi_A(t)|^2 |\psi_B(s)|^2 \right). \label{nonuniformF1bound} \end{align} Decreasing the fiber length, $L$, at constant nonlinearity strength, $\eta$, mitigates the phase-noise fidelity degradation in $F_0$. So long as $L$ satisfies the uniform phase-shift condition given in Eq.~(\ref{eq:uniform-phase-2}), $F_1$ will equal $F_0$, but once $L$ violates that condition we encounter a tradeoff for $F_1^\text{max}$ in Eq.~(\ref{nonuniformF1bound}): the phase-noise factor, $\exp[-(\eta L/4\ensuremath{\text{\foreignlanguage{greek}{p}}}\xspace)\int \dd{\omega} H_\text{im}(\omega)]$, decreases with further decreases in $L$, but the factor it multiplies will be greater than the $-1$ value it had when the phase shift was uniform. In Figs.~\ref{fig:f1-delta} and \ref{fig:f1-fast} we explore that tradeoff. \begin{figure} \centering \includegraphics[scale=0.4]{fiber-f1-delta.eps} \caption{\label{fig:f1-delta} (Color) Heat map of $F_1^\text{max}$ versus $\eta u$ and $L/u$ for Dirac-delta pulses with $t_d=L/2u$.} \end{figure} \begin{figure} \centering \includegraphics[scale=0.4]{fiber-f1-fast-response-fine.eps} \caption{\label{fig:f1-fast} (Color) Heat map of $F_1^\text{max}$ versus $\eta u$ and $L/u$ for Gaussian pulses with $t_\psi = 3\,\text{ps}$ and $t_d=L/2u$.} \end{figure} Figure~\ref{fig:f1-delta} shows a heat map of $F_1^\text{max}$ as $\eta u$ and $L/u$ are varied. Here we have assumed the extreme slow-response case of Dirac-delta pulses, and taken $t_d=L/2u$, so that the walk-off between the pulses is symmetric. It turns out that $F_1^\text{max}$ peaks at approximately 78.6\% when $\eta u \approx 4.25$ and $L/u\approx 16\,\text{fs}$. Although this peak value exceeds the 67.1\% $F_\text{max}$ value for fused-silica fiber, it is not very high and is lower than the optimum we gave earlier for the single-resonance, two-pole response function under uniform-phase-shift conditions. Figure~\ref{fig:f1-fast} shows a similar $F_1^\text{max}$ heat map for 3-ps-duration Gaussian pulses, i.e., $\psi_A(t) = \text{e}^{-2 t^2/t_\psi^2}/(\ensuremath{\text{\foreignlanguage{greek}{p}}}\xspace t_\psi^2/4)^{1/4}$ with $t_\psi = 3\,\text{ps}$. Here we see that the fidelity is abysmal, and our numerical calculation does not yield an $F_1^\text{max} > 2/3$. As expected, the uniform-phase-shift conditions are important here---causing the fidelity to be tightly bounded by the phase-noise alone---because operation is well into the fast-response regime. Taken together, our fidelity bounds for the theoretical and measured response functions permit us to confidently say that XPM in silica-core fiber cannot promise a high-fidelity $\ensuremath{\text{\foreignlanguage{greek}{p}}}\xspace$-radian {\sc cphase} gate, even under exceedingly idealistic assumptions.\\ \section{Conclusions} We have presented a continuous-time, quantum theory for cross-phase modulation with differing group velocities and have provided a framework for evaluating the fidelity of using quantum XPM to construct a {\sc cphase} gate. We found that perfect fidelity is impossible, even in theory, owing to causality-induced phase noise associated with Raman scattering in fused-silica fiber. For a reasonable theoretical response function and the experimentally-measured response function of silica-core fiber, we found that XPM will not support a high-fidelity {\sc cphase} gate, even under a collection of strictly favorable assumptions. In particular, our analysis ignores loss, dispersion, and self-phase modulation. Loss is especially pernicious, considering the length of fused-silica fiber needed for a single-photon pulse to create a $\ensuremath{\text{\foreignlanguage{greek}{p}}}\xspace$-radian phase shift on another such pulse. It is worth noting that the silica-core fiber response function we studied is that for co-polarized pulses. The response function for orthogonally-polarized pulses is much faster than---and 1/3 the strength of---its co-polarized counterpart, owing to its being mediated by an electronic interaction, as opposed to the Raman effect that is responsible for co-polarized XPM. We are not aware of any experimental characterization of the co-polarized response function. Nevertheless, the results in this paper suggest that it too will likely lead to low fidelity, so long as it is non-instantaneous, if for no other reason than that its extreme speed will force operation in the less favorable fast-response regime. Some final comments are now germane with respect to what potential {\sc cphase} gates are \em not\/\rm, as yet, precluded by our analysis. First, our results do not apply to XPM contained within a larger interaction system, such as a cavity. Some recent results have suggested that cavity-like systems may support a high-fidelity {\sc cphase} gate, despite noise \cite{PhysRevA.87.042325,PhysRevLett.110.223901}. To date, however, no one has studied the {\sc cphase}-gate fidelity afforded by cross-Kerr effect XPM within a cavity. Finally, it is unclear to what extent if at all our results apply to dark-state-polariton XPM in electromagnetically-induced transparency (EIT). EIT theories usually assume an instantaneous interaction, which is sometimes taken to be nonlocal \cite{PhysRevA.83.053826}. In the physical world, however, a phenomenon is rarely truly instantaneous, regardless of how good an approximation that may be for various working theories. Our work suggests that phase noise may be an issue for EIT if the response function is not truly instantaneous. That aside, recent work has shown that even instantaneous, nonlocal XPM is subject to the same fidelity-degrading phase noise, with limited exceptions \cite{Marzlin:10}. Together with the fact that EIT involves Raman interactions \cite{PhysRevLett.84.5094}, which are ultimately responsible for phase noise in co-polarized fiber XPM, this suggests that these systems might have to contend with the sort of fidelity issues presented here. Beyond that, other work has quantified additional fidelity-limiting issues, which may be present in continuous-time XPM, that seem likely to affect EIT systems \cite{banacloche, he1, he2}. \section*{ACKNOWLEDGMENTS} This research was supported by the DARPA Quantum Entanglement Science and Technology (QuEST) program and the NSF IGERT program Interdisciplinary Quantum Information Science and Engineering (iQuISE). G.~P.~Agrawal graciously provided the Raman response data originally collected by R.~H.~Stolen. \newcommand{\noopsort}[1]{} \newcommand{\printfirst}[2]{#1} \newcommand{\singleletter}[1]{#1} \newcommand{\switchargs}[2]{#2#1}
\section{Introduction} Physiological signals are signals that are measured from sensors that are either placed on or implanted into the body. Such physiological signals include those obtained using electromyography (EMG), electrocardiography (ECG), electroencephalography (EEG), photoplethysmography (PPG), and ballistocardiography (BCG). The processing and interpretation of such signals is challenging due to a number of different factors. For example, it is often difficult to obtain high-fidelity physiological signals due to noise, resulting in low signal-to-noise ratio (SNR). Traditionally, signal averaging and linear filters such as band-reject and band-pass filters have been used to process such physiological signals to suppress noise; however, such approaches have also been shown to result in signal degradation~\cite{Christov,Mewette}. As such, more advanced methods for handling such physiological signals are desired. Multi-scale decomposition has become an invaluable tool for the processing of physiological signals. In multi-scale decomposition, a signal is decomposed into a set of signals, each characterizing information about the original signal at a different scale. A common signal processing task that multi-scale decomposition has shown to provide significant benefits is noise suppression, based on the notion that the information pertaining to the noise component would be largely characterized by certain scales that are separate from the scales characterizing the desired signal. Much of literature in multi-scale decomposition for physiological signal processing has focused on scale-space theory~\cite{Melo08,Jager,Witkin83,Koenderink,Perona90,Gilboa08,scale1,scale2} and wavelet transforms~\cite{Phinyomark,Kestler,Hussain,Sobahi,Popescu,Chouakri,Agante,Donoho1,Donoho2}, with some investigations also conducted using methods such as empirical mode decomposition~\cite{Kopsinis,Weng,Velasco}. In scale-space theory~\cite{Witkin83}, a signal $f(t)$ is decomposed into a single-parameter family of $n$ signals, denoted by $L$, with a progressive decrease in fine scale signal information between successive scales: \begin{equation} L = \{l_{j}(t) | 0 \leq j \leq n-1\}, \end{equation} \noindent where $t$ denotes time, $j$ denotes scale, $l_{j}(t)$ is the signal at the ${j}^{\rm th}$ scale, and $l_0(t)=f(t)$. By decomposing a signal into a set of signals with a progressive decrease in fine scale signal information between successive scales, one can then analyze signals at coarser scales without the influence of fine scale signal information such as that pertaining to noise, which is mainly characterized at the finer scales. As such, one can utilize scale space theory to suppress noise in a signal by perform scale space decomposition on the signal and then treating one of signals at a coarser scale as the noise-suppressed signal. However, there are several limitations to the use of scale space theory for physiological signal processing pertaining to noise suppression. First, noise suppression using scale space theory requires the careful selection of which scale represents the noise-suppressed signal, which can be challenging. Second, noise suppression using scale space theory does not facilitate for fine-grained noise suppression at the individual scales, which limits its overall flexibility in striking a balance between noise suppression and signal structural preservation. In wavelet decomposition~\cite{Mallat89,Daubechies92}, a signal $f(t)$ is decomposed into a set of wavelet coefficients $c_{j,k}(t)$ obtained using a wavelet transform $W$: \begin{equation} c_{j,k}(t)=W_{\psi,f}(a,b)(t) \end{equation} \noindent where $\psi$ is the wavelet, $a=2^{-j}$ is the dyadic dilation, and $b=k2^{-j}$ is the dyadic position. Wavelet transforms has a number of advantages for the purpose of physiological signal processing, particularly pertaining to noise suppression. First, as signal information at different scales are better separated in the wavelet domain (i.e., signal information at one scale is not contained in another scale), this facilitates fine-grained noise suppression at the individual scales to strike a balance between noise suppression and signal structural preservation. Second, scale selection when performing noise suppression using wavelet transforms is less critical than that for noise suppression using scale space theory, since all scales are considered in noise suppression using wavelet transforms as opposed to a single scale selection with scale space theory. One limitation worth noting pertaining to signal processing using wavelet transforms, particularly pertaining to noise suppression, is that signals processed using wavelet transforms can exhibit oscillation artifacts related to wavelet basis functions used in the wavelet transform, which is particular noticeable when dealing with low SNR scenarios. Therefore, given some of the limitations with both scale space theory and wavelet transforms when used for physiological signal processing, one is motivated to explore alternative approaches that can address these limitations. Here, we take a different approach by exploring a Bayesian perspective to multi-scale signal decomposition. In this perspective, a signal is viewed as an amalgamation of a number of signals, each characterizing unique signal information at a different scale with different statistical characteristics. Taking such a perspective to the problem of multi-scale signal decomposition has a number of advantages. First, like the wavelet transform, since signal information at one scale is not contained in another scale, it allows us to achieve the benefits of taking better advantage of fine-grained noise suppression at the individual scales to strike a balance between noise suppression and signal structural preservation. Second, since signals are decomposed based on their statistical characteristics as opposed to a set of deterministic basis functions, signals processed using this approach would not exhibit the types of basis-related artifacts associated with the use of wavelet transforms. Motivated by this, in this study, we investigate the feasibility of utilizing a new Bayesian-based method for multi-scale signal decomposition called Bayesian Residual Transform (BRT) for the purpose of physiological signal processing. This paper is organized as follows. First, the methodology behind the proposed Bayesian Residual Transform is described in Section~\ref{methods}. The experimental setup for evaluating the feasibility of using the BRT for suppressing noise in physiological signals via signal-to-noise ratio (SNR) analysis using electrocardiography (ECG) signals is described in Section~\ref{setup}. The experimental results and discussion is presented in Section~\ref{results}, and conclusions are drawn and future work discussed in Section~\ref{conclusions}. \section{Bayesian Residual Transform} \label{methods} \begin{figure*}[tp] \centering \includegraphics[width=0.6\linewidth]{fig1} \caption{Bayesian Residual Transform framework. \textbf{(a)} forward BRT. \textbf{(b)} inverse BRT.} \label{fig1} \end{figure*} A full derivation of the proposed Bayesian Residual Transform (BRT) can be described as follows. In the BRT, a signal $f(t)$ is modeled as the summation of $n$ residual signals, each characterizing signal information from the signal at increasingly coarse scales: \begin{equation} f(t) = \sum_{i=1}^{n}r_i(t) = f_{\sum,1}(t), \label{sumofprocesses} \end{equation} \noindent where $f_{\sum,j}(t)$ denote a signal representing the summation of all residual signals at scales $[j,n]$: \begin{equation} f_{\sum,j}(t) = \sum_{i=j}^{n}r_i(t), \label{fsum} \end{equation} \noindent and $r_{i}=\{r_{i}(t) | t \in T\}$ is a residual signal characterizing the signal information at the $i^{\rm th}$ scale with different statistical characteristics. The residual signals at the lower scales contain fine-grained signal characteristics of the signal, while the residual signals at the higher scales contain coarse-grained signal characteristics of the signal (e.g., $r_1(t)$ characterizes the finest-grained signal characteristics, while $r_n(t)$ characterizes the coarsest-grained signal characteristics). As such, each residual signal contains unique information about the scale corresponding to a particular scale that the other residual signals do not contain. Therefore, the goal of the BRT (denoted by the function $B$) is to decompose a signal $f(t)$ into the set of $n$ residual signals $r_1(t), r_2(t), \ldots, r_n(t)$: \begin{equation} \{r_1(t), r_2(t), \ldots, r_n(t)\}=B(f(t)). \end{equation} \noindent Determining the set of residual signals characterizing the signal information at the different scales and whose sum is equal to $f(t)$ (i.e., Eq.~\ref{sumofprocesses}) is a highly challenging problem, and as such with the BRT we wish to introduce a deep cascading framework to solve this problem in a more tractable manner, where a residual signal at a particular scale is computed based on computations performed at a previous scale. Let us first rewrite Eq.~\ref{sumofprocesses} as follows: \begin{equation} f_{\sum,1}(t) = f_{\sum,2}(t) + r_1(t). \label{process2} \end{equation} \noindent It can be observed from Eq.~\ref{process2} that the residual signal $r_1(t)$ can be treated as the residual between the summation of all residual signals at scales $[1,n]$ and the summation of all residual signals at scales $[2,n]$. Hence, one can treat this as an inverse problem of estimating $f_{\sum,2}(t)$ given $f_{\sum,1}(t)$, with the analytical solution given by the conditional expectation $E(f_{\sum,2}(t)|f_{\sum,1}(t))$~\cite{Fieguth} (the quantification of the conditional expectation will be explained in more detail in a later section discussing the realization of the BRT via kernel regression). Therefore, given ${\hat {f}}_{\sum,2}(t)=E(f_{\sum,2}(t)|f_{\sum,1}(t))$, one can substitute ${\hat {f}}_{\sum,2}(t)$ for ${f}_{\sum,2}(t)$ in Eq.~\ref{process2} and rearrange the terms to obtain $r_1(t)$ as: \begin{equation} r_1(t) = f_{\sum,1}(t) - {\hat {f}}_{\sum,2}(t). \label{process5} \end{equation} \noindent Given ${\hat {f}}_{\sum,2}(t)$, which is computed to obtain $r_1(t)$, we can express the relationship between $r_2(t)$ and ${\hat {f}}_{\sum,2}(t)$ in a similar manner to Eq.~\ref{process2} as: \begin{equation} {\hat {f}}_{\sum,2}(t) = f_{\sum,3}(t) + r_2(t), \label{process5c} \end{equation} \noindent which can similarly be treated as an inverse problem of estimating $f_{\sum,3}(t)$ given ${\hat {f}}_{\sum,2}(t)$, with the analytical solution given by the conditional expectation $E(f_{\sum,3}(t)|{\hat {f}}_{\sum,2}(t))$. Therefore, given ${\hat {f}}_{\sum,3}(t)=E(f_{\sum,3}(t)|{\hat {f}}_{\sum,2}(t))$, one can express $r_2(t)$ as: \begin{equation} r_2(t) = {\hat {f}}_{\sum,2}(t) - {\hat {f}}_{\sum,3}(t). \label{process5c} \end{equation} \noindent Generalizing this, $r_j(t)$ at scale $j$, for $j<n$, can be obtained by \begin{equation} r_j(t) = {\hat {f}}_{\sum,j}(t) - {\hat {f}}_{\sum,j+1}(t), \label{estj} \end{equation} \noindent where \begin{equation} {\hat {f}}_{\sum,j+1}(t)=E(f_{\sum,j+1}(t)|{\hat {f}}_{\sum,j}(t)). \label{condexpectation} \end{equation} \noindent The last residual signal is computed as $r_n(t)={\hat {f}}_{\sum,n}(t)$ to conform with the form expressed in Eq.~\ref{sumofprocesses}. Hence, given Eq.~\ref{estj}, we have a deep cascading framework for the BRT where we can obtain the residual signal at scale $j$ (i.e., $r_j(t)$) given the previously computed ${\hat {f}}_{\sum,j}(t)$. Furthermore, since the residual signal at scale $j-1$ (i.e., $r_{j-1}(t)$) is not involved in the computation of the residual signal at scale $j$ (i.e., $r_j(t)$) (only ${\hat {f}}_{\sum,j}(t)$ obtained from previous cascading step is), the information contained within $r_{j-1}(t)$ is not contained within $r_{j}(t)$. As such, as scale $j$ increases, the signal information contained in ${\hat {f}}_{\sum,j}(t)$ becomes coarser and coarser, which results in residual signals $r_{j}(t)$ characterizing coarser and coarser signal information as scale increases. Based on Eq.~\ref{estj}, the deep cascading framework for the forward Bayesian Residual Transform (BRT) is illustrated in Fig.~\ref{fig1}a. Due to the condition of the summation of residual signals at all scales being equal to signal $f(t)$ (Eq.~\ref{sumofprocesses}), the inverse BRT is simply the summation of all residual signals $r_1(t), r_2(t), \ldots, r_n(t)$: \begin{equation} f(t) = B^{-1}(r_1(t), r_2(t), \ldots, r_n(t))=\sum_{i=1}^{n}r_i(t). \label{inversetransform} \end{equation} \noindent The inverse Bayesian Residual Transform (inverse BRT) procedure is illustrated in Fig.~\ref{fig1}b. \subsection{Realization of Bayesian Residual Transform via Kernel Regression} In this study, we implement a realization of the BRT using a kernel regression strategy, which can be described as follows. At each iteration $j$, we compute $E(f_{\sum,j+1}(t)|{\hat {f}}_{\sum,j}(t))$ (Eq.~\ref{condexpectation}) based on nonparametric Nadaraya-Watson kernel regression~\cite{Nadaraya,Watson} using a kernel function $K_j$. Here, we employ the following Gaussian kernel function $K_j$: \begin{equation} K_j({\hat {f}}_{\sum,j}(t)-{\hat {f}}_{\sum,j}(t_i))=e^{-\frac{1}{\lambda_j^2}({\hat {f}}_{\sum,j}(t)-{\hat {f}}_{\sum,j}(t_i))^2} \label{kj} \end{equation} \noindent Finally, the residual signal at scale $n$ (i.e., $r_n(t)$) can be set as $E(f_{\sum,n}(t)|{\hat {f}}_{\sum,n-1}(t))$, which is computed at the step where $r_{n-1}(t)$ is computed. By setting $r_n(t)=E(f_{\sum,n}(t)|{\hat {f}}_{\sum,n-1}(t))$, the condition of the summation of signal decompositions at all scales being equal to signal $f(t)$ (i.e., Eq.~\ref{sumofprocesses}) is satisfied. A step-by-step summary of the realization of BRT via kernel regression is shown in Algorithm~\ref{alg1}. A step-by-step summary of the inverse BRT is shown in Algorithm~\ref{alg2}. \subsection{Noise suppression} \label{noisesuppression} In this study, we wish to illustrate the feasibility of utilizing the BRT for processing physiological signals through the task of noise suppression. As such, we first establish a simple approach to noise suppression of signals using the BRT for illustrative purposes. The noise suppression method chosen for this study is based around the idea that the observed noisy signal $f(t)$ is formed as a summation of the desired noise-free signal $f'(t)$ and an additive noise source. Suppose that we have the true noise-free signal $f'(t)$ and we decompose it using the BRT into a series of residual signals $r_1(t), r_2(t), \ldots, r_n(t)$, where each of the residual signals characterize only information from the noise-free signal at a particular scale. Much of the information at each scale that characterizes the noise-free signal $f'(t)$ would be concentrated within only a few of the locations in each of the residual signals. What this means is that much of the information content related to $f'(t)$ is primarily concentrated within just a few locations at each scale. If we were to decompose the noisy signal $f(t)$ using the BRT in a similar fashion, the locations of the residual signal at each scale that would otherwise have negligible information content associated with $f'(t)$ would now have low but not negligible information content that characterizes the noise source. Motivated by this, we employ a noise thresholding strategy where we only keep information from locations with information content greater than the noise information content level $\theta$ at each scale. \begin{algorithm}[h] \caption{Step-by-step summary for Bayesian Residual Transform via Kernel Regression} \begin{algorithmic} \REQUIRE~~\\{\STATE A signal $f(t)$} \STATE parameters initialization: $\lambda_1,\ldots,\lambda_{n-1}, n$ \ENSURE~~\\{\STATE residual signals $r_1(t), r_2(t), \ldots, r_n(t)$}\\~\\ \STATE $j=1$; \STATE $\hat f_{\sum,1}(t)=f(t)$; \WHILE {$(j < n )$} \STATE Compute $\hat f_{\sum,j+1}(t)=E(f_{\sum,j+1}(t)|{\hat {f}}_{\sum,j}(t))$ based on kernel regression with $K_j$ $\gets$ Eq. (\ref{kj})\ \STATE Compute $r_j(t) = {\hat {f}}_{\sum,j}(t) - \hat f_{\sum,j+1}(t) \gets$ Eq. (\ref{estj})\ \STATE $j=j+1$; \ENDWHILE \STATE $r_n(t) = \hat f_{\sum,j}(t)$ \end{algorithmic} \label{alg1} \end{algorithm} Motivated by this, the noise thresholding strategy employed in this study can be described as follows. We first perform the forward BRT on the signal $f(t)$ to obtain $n$ residual signals characterizing signal information at different scales ($r_1(t), r_2(t), \ldots, r_n(t)$). Since the noise information content level at each scale is not known, we employ the seminal noise level estimation method proposed by Donoho~\cite{Donoho2} to determine the noise threshold $\theta$ at each scale, which can be described as follows. At scale $j$, we estimate the noise threshold $\theta_j$ at each scale $j$ using the noise-adaptive scale estimate, which can be expressed by: \begin{equation} \theta_j = MAD(r_j) / {\Phi^{-1}(3/4)}, \label{noiseest2} \end{equation} \noindent where $MAD$ is the median absolute deviation and $\Phi^{-1}$ is the normal inverse cumulative distribution function. Based on $\theta_j$, noise thresholding is achieved to obtain noise-suppressed residual signal $r'_j(t)$ by: \begin{align} \hspace*{0.25in} r'_j(t)&=\left\{\begin{array}{ccccc} 0 & \textup{if } & |r_j(t)| < & \theta_j \\ r_j(t) & \textup{if } & & otherwise & \\ \end{array} \right. \label{eqn-2} \end{align} \begin{figure*}[tp] \centering \includegraphics[width=0.7\linewidth]{fig2} \caption{Example of multi-scale signal decomposition using the BRT. \textbf{(a)} Baseline periodic test signal. \textbf{(b)} Noisy input signal with zero-mean Gaussian noise. \textbf{(c)-(h)} Signal decompositions using the BRT at different scales. It can be observed that the noise process contaminating the test signal is well characterized in the decompositions at the lower (finer) scales (scales 1 to 3), while the structural characteristics of the test signal is well characterized in the decompositions at the higher (coarser) scales (scales 4 to 6). } \label{fig2} \end{figure*} \noindent Finally, the inverse BRT (Eq.~\ref{inversetransform}) is performed on the set of $n$ noise-suppressed residual signals at the different scales ($r'_1(t), r'_2(t), \ldots, r'_n(t)$) to produce the noise-suppressed signal $f'(t)$. A step-by-step summary of the noise suppression method using the BRT is shown in Algorithm~\ref{alg3}.\\ \begin{algorithm}[h] \caption{Step-by-step summary for inverse Bayesian Residual Transform} \begin{algorithmic} \REQUIRE~~\\{\STATE residual signals $r_1(t), r_2(t), \ldots, r_n(t)$} \STATE parameters initialization: $n$ \ENSURE~~\\{\STATE A signal $f(t)$}\\~\\ \STATE $j=1$; \STATE $f(t)=0$; \WHILE {$(j \leq n )$} \STATE $f(t)=f(t)+r_j(t)$; \STATE $j=j+1$; \ENDWHILE \end{algorithmic} \label{alg2} \end{algorithm} \begin{algorithm}[h] \caption{Step-by-step summary for noise suppression using Bayesian Residual Transform} \begin{algorithmic} \REQUIRE~~\\{\STATE A noisy signal $f(t)$} \STATE parameters initialization: $n$ \ENSURE~~\\{\STATE A noise-suppressed signal $f'(t)$}\\~\\ \STATE Perform the BRT on $f(t)$ to obtain residual signals $r_1(t), r_2(t), \ldots, r_n(t)$. $\gets$ Algorithm~\ref{alg1} \STATE $j=1$; \WHILE {$(j \leq n )$} \STATE Compute noise threshold $\theta_j$ $\gets$ Eq.~\ref{noiseest2}; \STATE Compute noise-suppressed residual signal $r'_j(t)$ via threshold using $\theta_j$ $\gets$ Eq.~\ref{eqn-2}; \STATE $j=j+1$; \ENDWHILE \STATE Perform inverse BRT on $r'_1(t), r'_2(t), \ldots, r'_n(t)$ to obtain noise-suppressed signal $f'(t)$ $\gets$ Algorithm~\ref{alg2} \end{algorithmic} \label{alg3} \end{algorithm} \begin{figure*}[tp] \centering \includegraphics[width=0.8\linewidth]{fig3} \caption{Example of multi-scale signal decomposition using the BRT. \textbf{(a)} Baseline piece-wise regular test signal. \textbf{(b)} Noisy input signal with zero-mean Gaussian noise. \textbf{(c)-(h)} Signal decompositions using the BRT at different scales. It can be observed that, as with the periodic signal example, the noise process contaminating the signal is well characterized in the decompositions at the lower (finer) scales (scales 1 to 2), while the structural characteristics of the test signal is well characterized in the decompositions at the higher (coarser) scales (scales 3 to 6). Furthermore, more noticeable here than in the periodic example, it can be seen that that the decomposition at each scale exhibits good signal structural localization.} \label{fig3} \end{figure*} \section{Experimental setup} \label{setup} In this study, to illustrate the feasibility of utilizing the BRT for processing physiological signals, we performed a SNR analysis using electrocardiography (ECG) signals to study the performance of the BRT for the task of noise suppression. ECG signals from the MIT-BIH Normal Sinus Rhythm Database~\cite{Sinus} were used in this study to perform the SNR analysis. This database consists of 18 ECG recordings (recorded at a sampling rate of 128 Hz) of subjects conducted at the Arrhythmia Laboratory in the Beth Israel Deaconess Medical Center. The subjects were found to have no significant arrhythmias. A total of 18 low-noise segments of 10 seconds was extracted, one from each recording, based on visual inspection to act as the baseline signals for evaluation. To study noise suppression performance at different SNR levels, each of the 18 baseline signals were contaminated by white Gaussian noise to produce noisy signals with SNR ranging from 12 dB to 2.5 dB (with 20 different noisy signals at each SNR), resulting in 3960 different signal perturbations used in the analysis. For comparison purposes, wavelet denoising methods with the following shrinkage rules were also used: i) Stein's Unbiased Risk (SURE)~\cite{Donoho3}, ii) Heuristic SURE (HSURE)~\cite{matlab}, iii) Universal (UNI)~\cite{matlab}, and iv) Minimax (MINIMAX)~\cite{Donoho2}. Each of the methods uses their corresponding noise threshold and shrinkage rules in the original works. To quantitative evaluate noise suppression performance, we compute the SNR improvement as follows~\cite{Akhbari}: \begin{equation} SNRI = 10\log\left(\frac{\sum_t(f(t) - f_b(t))^2}{\sum_t(f'(t) - f_b(t))^2}\right) \end{equation} \noindent where $f(t)$, $f_b(t)$, and $f'(t)$ are the noisy, baseline, and noise-suppressed signals obtained using a noise suppression method, respectively.\\ ~\\ To study the effect of the number of scales $n$ on noise suppression performance, the same SNR analysis is performed as described above for $n=\{2,3,4,5,6\}$.\\ \subsection{Implementation details} The BRT is implemented in MATLAB (The MathWorks, Inc.), with the nonparametric conditional expectation estimates implemented in C++ and compiled as a dynamically linked MATLAB Executable (MEX) to improve computational speed. The only free parameters of the implemented realization of the BRT are the standard deviations used to model the residual signals (e.g., $\lambda$), the number of scales $n$, and time window size, which can be adjusted by the user to find a tradeoff between noise suppression quality and computational costs. For the SNR analysis of ECG signals, $\lambda$ is set equally for all scales to the standard deviation of $f(t)$ for simplicity, $n$ is set at 6 scales, and the time window size is set to $0.1s$. For this configuration, the current implemented realization of the BRT can process a 1028-sample signal in $<$1 second on an Intel(R) Core(TM) i5-3317U CPU at 1.70GHz CPU. For the wavelet-based methods tested (SURE, HSURE, UNI, and MINIMAX), as implemented in MATLAB (The MathWorks, Inc.), soft thresholding with the Coiflet3 mother wavelet at 6 scales and single level rescaling was used as it was found to provide superior results for ECG noise suppression~\cite{Sameni}. Each of the methods uses their corresponding noise threshold and shrinkage rules as specified in the original works. \section{Experimental Results} \label{results} To illustrate the feasibility of utilizing the BRT for processing physiological signals, such as for the task of noise suppression, we first performed the BRT on two test signals: i) a noisy periodic test signal, and ii) a noisy piece-wise regular test signal. The multi-scale signal decomposition using the BRT on a noisy periodic test signal is shown in Fig.~\ref{fig2}. Here, a baseline test signal (Fig.~\ref{fig2}a) is contaminated by a zero-mean Gaussian noise process to produce a noisy signal (Fig.~\ref{fig2}b) and then decomposed using the BRT at different scales (Figs.~\ref{fig2}c-h). It can be observed that the noise process contaminating the signal is well characterized in the decompositions at the lower (finer) scales (scales 1 to 3), while the structural characteristics of the test signal is well characterized in the decompositions at the higher (coarser) scales (scales 4 to 6). \begin{figure*}[tp] \centering \includegraphics[width=0.9\linewidth]{fig4} \caption{Application of the BRT on ECG signals. \textbf{(a)} A plot of the mean SNR improvement vs. the different input SNRs ranging from 12 dB to 2.5 dB for the MIT-BIH Normal Sinus Rhythm Database for the tested methods. Noise-suppression method using the BRT provided strong SNR improvements across all SNRs, with performance comparable to SURE and higher than the other 3 tested methods. \textbf{(b)} A plot of the mean SNR improvement vs. the different input SNRs ranging from 12 dB to 2.5 dB for the method using the BRT with different number of scales $n$. \textbf{(c)} A plot of the mean SNR improvement vs. the different input SNRs ranging from 12 dB to 2.5 dB for the method using the BRT with different multiples of the standard deviation (SD) for $\lambda$.} \label{fig4} \end{figure*} The multi-scale signal decomposition using the BRT on a noisy piece-wise regular test signal (generated using~\cite{tp}) is shown in Fig.~\ref{fig3}. As with the previous example, a baseline test signal (Fig.~\ref{fig3}a) is contaminated by a zero-mean Gaussian noise process to produce a noisy signal (Fig.~\ref{fig3}b) and then decomposed using the BRT at different scales (Figs.~\ref{fig3}c-h). It can be observed that, as with the periodic signal example, the noise process contaminating the signal is well characterized in the decompositions at the lower (finer) scales (scales 1 to 2), while the structural characteristics of the test signal is well characterized in the decompositions at the higher (coarser) scales (scales 3 to 6). Furthermore, more noticeable here than in the periodic signal example, it can be seen that that the decomposition at each scale exhibits good signal structural localization. Therefore, given the ability of the BRT to decouple the noise process from the true signal into different scales, as illustrated in both the periodic and piece-wise regular test signals, the BRT has the potential to be useful for performing noise suppression on signals while preserving inherent signal characteristics. In this study, to illustrate the feasibility of utilizing the BRT for processing physiological signals, we introduced a simple thresholding approach to noise suppression using the BRT for illustrative purposes (\textbf{see Section~\ref{noisesuppression}}). We then performed a quantitative SNR analysis using electrocardiography (ECG) signals from the MIT-BIH Normal Sinus Rhythm Database~\cite{Sinus} to study the performance of the BRT for the task of noise suppression, where the SNR improvement (\textbf{see Section~\ref{noisesuppression} for formulation}). A plot of the mean SNR improvement of the tested methods vs. the different input SNRs ranging from 12 dB to 2.5 dB is shown in Fig.~\ref{fig4}a. It can be observed that the noise-suppression method using the BRT provided strong SNR improvements across all SNRs, comparable to SURE and higher than the other 3 tested methods. It can also be observed that the UNI method consistently achieved SNR improvements below 0 dB. This is primarily due to the tendency to overestimate the noise level, resulting in signal oversmoothing and thus producing a noise-suppressed signal that is less similar to the baseline signal than the actual noisy signal. It can also be observed that the SNR improvement increases as the SNR of the input noisy signal decreases, which indicates that greater benefits are obtained through the use of noise suppression methods in low signal SNR scenarios. To study the effect of the number of scales $n$ on noise suppression performance, a plot of the mean SNR improvement vs. the different input SNRs ranging from 12 dB to 2.5 dB for the method using the BRT with a range of different number of scales $n$ is shown in Fig.~\ref{fig4}b. It can be observed that a significant gain in SNR improvement exists going from $n=2$ to $n=3$, with smaller SNR improvement gains from $n=3$ all the way to $n=6$. Furthermore, it can be observed that the SNR improvement gains from increasing the number of scales become smaller and smaller as the input SNR decreases, with the SNR improvement for $n=3$ to $n=6$ being approximately the same when the input SNR is 2.5 dB. Therefore, this indicates that the effect of selecting the number of scales on noise suppression performance can be significant and thus a balance between SNR improvement and the computational complexity of the BRT (which grows linearly with the number of scales) is necessary, particularly given the SNR of the noisy signal. To study the effect of the standard deviation (SD) used for $\lambda$ on noise suppression performance, a plot of the mean SNR improvement vs. the different input SNRs ranging from 12 dB to 2.5 dB for the method using the BRT with a range of different multiples of SD used for $\lambda$ is shown in Fig.~\ref{fig4}c. It can be observed that a significant gain in SNR improvement exists going from $0.5SD$ to $1SD$, with a significant drop in SNR improvements going from $1SD$ to $2SD$. Furthermore, it can be observed that there are noticeable SNR improvement gains going from $0.5SD$ to $2SD$ that grows larger as the input SNR decreases. Therefore, this indicates that the effect of selecting $\lambda$ on noise suppression performance can be significant, and careful selection may be important when dealing with different types of signals. For the signals tested here, it was found that $1SD$ provided the strongest results. Typical results of noise-suppressed signals produced by the method using the BRT are shown in Fig.~\ref{fig5}b and Fig.~\ref{fig5}e (corresponding to two different 12 dB noisy input signals shown in Fig.~\ref{fig5}a and Fig.~\ref{fig5}d, respectively). Visually, it can be seen that the BRT was effectively used to produce signals with significantly reduced noise artifacts while preserving signal characteristics. Results in this study show that it is feasible to utilize the BRT for processing physiological signals for tasks such as noise suppression. \begin{figure*}[tp] \centering \includegraphics[width=0.8\linewidth]{fig5} \caption{Application of the BRT on ECG signals. \textbf{(a)} Noisy input signal with SNR=12 dB, \textbf{(b)} noise-suppressed results using BRT for \textbf{a}, and \textbf{(c)} the corresponding original signal. \textbf{(d)} Another noisy input signal with SNR=12 dB, and \textbf{(e)} noise-suppressed results using BRT for \textbf{d}, and \textbf{(f)} the corresponding original signal. The results produced using the BRT has significantly reduced noise artifacts while the signal characteristics are preserved. } \label{fig5} \end{figure*} \section{Conclusion} \label{conclusions} In this study, the feasibility of employing a Bayesian-based approach to multi-scale signal decomposition introduced here as the Bayesian Residual Transform for use in the processing of physiological signals. The Bayesian Residual Transform decomposes a signal into a set of residual signals, each characterizing information from the signal at different scales and following a particular probability distribution. This allows information at different scales to be decoupled for the purpose of signal analysis and, for the purpose of noise suppression, allows for information pertaining to the noise process contaminating the signal to be separated from the rest of the signal characteristics. This trait is important for performing noise suppression on signals while preserving inherent signal characteristics. SNR analysis using a set of ECG signals from the MIT-BIH Normal Sinus Rhythm Database at different noise levels demonstrated that it is feasible to utilize the BRT for processing physiological signals for tasks such as noise suppression. Given the promising results, we aim in the future to investigate alternative adaptive thresholding schemes for the task of noise suppression in physiological signals characterized by nonstationary noise, so that one can better adapt to the nonstationary noise statistics embedded at different scales. Moving beyond low-level signal processing tasks such as noise suppression, we aim with our future work to investigate and devise methods for multi-scale analysis of a signal using the Bayesian Residual Transform, which could in turn lead to improved features for signal classification. Finally, we aim to investigate the extension and generalization of the Bayesian Residual Transform for dealing with high-dimensional physiological signals such as vectorcardiographs (VCG)~\cite{Sameni2}, and dealing with high-dimensional medical imaging signals from systems such as multiplexed optical high-coherence interferometry~\cite{Farnoud}, optical coherence tomography~\cite{OCT1,OCT2}, dermatological imaging~\cite{derm}, diffusion weighted magnetic resonance imaging (DWI)~\cite{Koh,Bihan,Shafiee}, microscopy~\cite{microscopy1,microscopy2}, dynamic contrast enhanced MRI (DCE-MRI), and correlated diffusion imaging~\cite{Wong,dualstage}. \section{Acknowledgment} This work was supported by the Natural Sciences and Engineering Research Council of Canada, Canada Research Chairs Program, and the Ontario Ministry of Research and Innovation. \ifCLASSOPTIONcaptionsoff \newpage \fi
\section{Introduction}\label{introduction} An important problem in quantitative risk management is to aggregate several individually studied types of risks into an overall position. Mathematically, this translates into studying the worst-case distribution tails of $\Psi(X)$, where $\Psi:\R^n\rightarrow\R$ is a given function that represents the risk (or underiability) of an outcome, and where $X$ is a random vector that takes values in $\R^n$ and whose distribution is only partially known. For example, one may only have information about the marginals of $X$ and possibly partial information about some of the moments. To solve such problems, duality is often exploited, as the dual may be easier to approach numerically or analytically \cite{puccetti1, puccetti2, puccetti3, ramachandran, popescu}. Being able to formulate a dual is also important in cases where the primal is approachable algorithmically, as solving the primal and dual problems jointly provides an approximation guarantee throughout the run of a solve: if the duality gap (the difference between the primal and dual objective values) falls below a chosen threshold relative to the primal objective, the algorithm can be stopped with a guarantee of approximating the optimum to a fixed precision that depends on the chosen threshold. This is a well-known technique in convex optimization, see e.g.\ \cite{borwein-lewis}. Although for some special cases of the marginal problem analytic solutions and powerful numerical heuristics exist \cite{puccettinew1, puccettinew2, puccettinew3, wang1, wang2}, these techniques do not apply when additional constraints are imposed to force the probability measures over which we maximize the risk to conform with empirical observations: In a typical case, the bulk of the empirical data may be contained in a region $D$ that can be approximated by an ellipsoid or the union of several (disjoint or overlapping) polyhedra. For a probability measure $\mu$ to be considered a reasonable explanation of the true distribution of (multi-dimensional) losses, one would require the probability mass contained in $D$ to lie in an empirically estimated confidence region, that is, $\ell\leq\mu(D)\leq u$ for some estimated bounds $\ell<u$. In such a situation, the derivation of robust risk aggregation bounds via dual problems remains a powerful and interesting approach. In this chapter we formulate a general optimization problem, which can be seen as a doubly infinite linear programming problem, and we show that the associated dual generalizes several well known special cases. We then apply this duality framework to a new class of risk management models we propose in Section \ref{bounds on integrals}. \section{A General Duality Relation}\label{duality} Let $(\Phi,\fF)$, $(\Gamma,\fG)$ and $(\Sigma,\fS)$ be complete measure spaces, and let $A:\,\Gamma\times\Phi\rightarrow\R$, $a:\,\Gamma\rightarrow\R$, $B:\,\Sigma\times\Phi\rightarrow\R$, $b:\,\Sigma\rightarrow\R$, and $c:\,\Phi\rightarrow\R$ be bounded measurable functions on these spaces and the corresponding product spaces. Let $\cM_{\fF}$, $\cM_{\fG}$ and $\cM_{\fS}$ be the set of signed measures with finite variation on $(\Phi,\fF)$, $(\Gamma,\fG)$ and $(\Sigma,\fS)$ respectively. We now consider the following pair of optimization problems over $\cM_{\fF}$ and $\cM_{\fG}\times\cM_{\fS}$ respectively, \begin{align*} \text{(P)}\quad\sup_{\calF\in\cM_{\fF}}\,&\int_{\Phi}c(x)\diff \calF(x)\\ \text{s.t. }&\int_{\Phi}A(y,x)\diff \calF(x)\leq a(y),\quad(y\in\Gamma),\\ &\int_{\Phi}B(z,x)\diff \calF(x)= b(z),\quad(z\in\Sigma),\\ &\calF\geq 0, \end{align*} and \begin{align*} \text{(D)}\quad\inf_{(\calG,\calS)\in\cM_{\fG}\times\cM_{\fS}} \,&\int_{\Gamma}a(y)\diff \calG(y)+\int_{\Sigma}b(z)\diff\calS(z),\\ \text{s.t. }&\int_{\Gamma}A(y,x)\diff\calG(y)+\int_{\Sigma} B(z,x)\diff\calS(z)\geq c(x),\quad(x\in\Phi),\\ &\calG\geq 0. \end{align*} We claim that the infinite-programming problems (P) and (D) are duals of each other. \begin{theorem}[Weak Duality]\label{weak duality} For every (P)-feasible measure $\calF$ and every (D)-feasible pair $(\calG,\calS)$ we have \begin{equation*} \int_{\Phi}c(x)\diff\calF(x)\leq \int_{\Gamma}a(y)\diff\calG(y)+ \int_{\Sigma}b(z)\diff\calS(z). \end{equation*} \end{theorem} \begin{proof} Using Fubini's Theorem, we have \begin{align*} \int_{\Phi}c(x)\diff\calF(x)&\leq \int_{\Gamma\times\Phi}A(y,x)\diff(\calG\times\calF)(y,x)+ \int_{\Sigma\times\Phi}B(z,x)\diff(\calS\times\calF)(z,x)\\ &\leq\int_{\Gamma}a(y)\diff\calG(y)+ \int_{\Sigma}b(z)\diff\calS(z). \end{align*} \end{proof} In various special cases, such as those discussed in Section \ref{special cases}, strong duality is known to hold subject to regularity assumptions, that is, the optimal values of (P) and (D) coincide. Another special case under which strong duality applies is when the measures $\calF$, $\calG$ and $\calS$ have densities in appropriate Hilbert spaces, see the forthcoming DPhil thesis of the second author \cite{sergey}. We remark that the quantifiers in the constraints can be weakened if the set of allowable measures is restricted. For example, if $\calG$ is restricted to lie in a set of measures that are absolutely continuous with respect to a fixed measure $\calG_0\in\cM_{\fG}$, then the quantifier $(y\in\Gamma)$ can be weakened to $(\calG_0\text{-almost all }y\in\Gamma)$. \section{Classical Examples}\label{special cases} Our general duality relation of Theorem \ref{weak duality} generalizes many classical duality results, of which we now point out a few examples. Let $p(x_1,\dots,x_k)$ be a function of $k$ arguments. Then we write \begin{equation*} 1_{\{x: p(x)\geq 0\}}:=1_{\{y: p(y)\geq 0\}}(x)=\begin{cases}1\quad&\text{if }p(x)\geq 0,\\ 0\quad&\text{otherwise}. \end{cases} \end{equation*} In other words, we write the argument $x$ of the indicator function directly into the set $\{y: p(y)\geq 0\}$ that defines the function, rather than using a separate set of variables $y$. This abuse of notation will make it easier to identify which inequality is satisfied by the arguments where the function $1_{\{y: p(y)\geq 0\}}(x)$ takes the value $1$. We start with the Moment Problem studied by Bertsimas \& Popescu \cite{popescu}, who considered generalized Chebychev inequalities of the form \begin{align*} \text{(P')}\quad\sup_{X}\;&\probability[r(X)\geq 0]\\ \text{s.t. }&\expect_{\mu}[X_1^{k_1}\dots X_n^{k_n}]=b_{k},\quad(k\in J),\\ &X\text{ a random vector taking values in }\R^n,\nonumber \end{align*} where $r:\R^n\rightarrow\R$ is a multivariate polynomial and $J\subset \N^n$ is a finite sets of multi-indices. In other words, some moments of $X$ are known. By choosing $\Phi=\R^n$, $\Gamma=\emptyset$, $\Sigma=J\cup\{0\}$, \begin{align*} &B(k,x)=x_1^{k_1}\dots x_n^{k_n},\quad b(k)=b_k,\quad(k\in J),\\ &B(0,x)=\1_{\R^n},\quad b(0)=1, \end{align*} and $c(x)=\1_{\{x:\,r(x))\geq 0\}}$, where we made use of the abuse of notation discussed above. Problem (P') becomes a special case of the primal problem considered in Section \ref{duality}, \begin{align*} \text{(P)}\quad\sup_{\calF}\,&\int_{\R^n}\1_{\{x:\,r(x)\geq 0\}}\diff\calF(x)\\ \text{s.t. }&\int_{\R^n}x_1^{k_1}\dots x_n^{k_n}\diff\calF(x)=b_{k},\quad(k\in J),\\ &\int_{\R^n}1\diff\calF(x)=1,\\ &\calF\geq 0. \end{align*} Our dual \begin{align*} \text{(D)}\quad\inf_{(z,z_0)\in\R^{|J|+1}}\,&\sum_{k\in J} z_k b_k + z_0\\ \text{s.t. }&\sum_{k\in J} z_k x_1^{k_1}\dots x_n^{k_n} + z_0\geq \1_{\{x: r(x)\geq 0\}}, \quad(x\in\R^n) \end{align*} is easily seen to be identical with the dual (D') identified by Bertsimas \& Popescu, \begin{align*} \text{(D')}\quad\inf_{(z,z_0)\in\R^{|J|+1}}\,&\sum_{k\in J} z_k b_k + z_0\\ \text{s.t. }&\forall\,x\in\R^n, r(x)\geq 0\Rightarrow \sum_{k\in J} z_k x_1^{k_1}\dots x_n^{k_n} + z_0-1\geq 0,\\ &\forall\,x\in\R^n, \sum_{k\in J} z_k x_1^{k_1}\dots x_n^{k_n} + z_0\geq 0. \end{align*} Note that since $\Gamma,\Sigma$ are finite, the constraints of (D') are polynomial copositivity constraints. The numerical solution of semi-infinite programming problems of this type can be approached via a nested hierarchy of semidefinite programming relaxations that yield better and better approximations to (D'). The highest level problem within this hierarchy is guaranteed to solve (D') exactly, although the corresponding SDP is of exponential size in the dimension $n$, in the degree of the polymomial $r$, and in $\max_{k\in J}(\sum_i k_i)$. For further details see \cite{popescu, parrilo, lasserre}, and Section \ref{piecewise polynomial} below. Next, we consider the Marginal Problem studied by R\"uschendorf \cite{ruschendorf1, ruschendorf2} and Ramachandran \& R\"uschendorf \cite{ramachandran}, \begin{equation*} \text{(P')}\quad\sup_{\calF\in\cM_{F_1,\dots,F_n}}\,\int_{\R^n}h(x)\diff \calF(x), \end{equation*} where $\cM_{F_1,\dots,F_n}$ is the set of probability measures on $\R^n$ whose marginals have the cdfs $F_i$ $(i=1,\dots,n)$. Problem (P') can easily be seen as a special case of the framework of Section \ref{duality} by setting $c(x)=h(x)$, $\Phi=\R^n$, $\Gamma=\emptyset$, $\Sigma=\N_n\times\R$, $B(i,z,x)=\1_{\{y:\,y_i\leq z\}}$ (using the abuse of notation discussed earlier), and $b_i(z)=F_i(z)$ $(i\in\N_n,\,z\in\R)$, \begin{align*} \text{(P)}\quad\sup_{\calF}\,&\int_{\R^n}h(x)\diff\calF(x)\\ \text{s.t. }&\int_{\R}\1_{\{x_i\leq z\}}\diff\calF(x) =F_i(z),\quad(z\in\R, i\in\N_n)\\ &\calF\geq 0. \end{align*} Taking the dual, we find \begin{align*} \text{(D)}\quad\inf_{\calS_1,\dots,\calS_n}\,& \sum_{i=1}^n\int_{\R}F_i(z)\diff\calS_i(z)\\ \text{s.t. }&\sum_{i=1}^n\int_{\R}\1_{\{x_i\leq z\}}\diff\calS_i(z) \geq h(x),\quad(x\in\R^n). \end{align*} The signed measures $\calS_i$ being of finite variation, the functions $S_i(z)=\calS((-\infty,z])$ and the limits $s_i=\lim_{z\rightarrow \infty}S_i(z)=\calS((-\infty,+\infty))$ are well defined and finite. Furthermore, using $\lim_{z\rightarrow-\infty}F_i(z)=0$ and $\lim_{z\rightarrow+\infty}F_i(z)=1$, we have \begin{align*} \sum_{i=1}^n\int_{\R}F_i(z)\diff\calS(z)&= \sum_{i=1}^n\left(F_i(z)S_i(z)|^{+\infty}_{-\infty}-\int_{\R}S_i(z)\diff F_i(z) \right)\\ &=\sum_{i=1}^n s_i - \sum_{i=1}^n\int_{\R}S_i(z)\diff F_i(z)\\ &=\sum_{i=1}^n\int_{\R}(s_i-S_i(z))\diff F_i(z), \end{align*} and likewise, \begin{equation*} \sum_{i=1}^n\int_{\R}\1_{\{x_i\leq z\}}\diff\calS_i(z)= \sum_{i=1}^n\int_{x_i}^{+\infty}1\diff\calS_i(z) =\sum_{i=1}^n (s_i - S_i(x_i)). \end{equation*} Writing $h_i(z)=s_i-S_i(z)$, (D) is therefore equivalent to \begin{align*} \text{(D')}\quad\inf_{h_1,\dots,h_n}\,&\sum_{i=1}^n\int_{\R}h_i(z)\diff F_i(z)\\ \text{s.t. }&\sum_{i=0}^n h_i(x_i)\geq h(x),\quad (x\in\R^n). \end{align*} This is the dual identified by Ramachandran \& R\"uschendorf \cite{ramachandran}. Due to the general form of the functions $h_i$, the infinite programming problem (D') is not directly usable in numerical computations. However, for specific $h(x)$, (D')-feasible functions $(h_1,\dots,h_n)$ can sometimes be constructed explicitly, yielding an upper bound on the optimal objective function value of (P') by virtue of Theorem \ref{weak duality}. Embrechts \& Puccetti \cite{puccetti1,puccetti2,puccetti3} used this approach to derive quantile bounds on $X_1+\dots+X_n$, where $X$ is a random vector with known marginals but unknown joint distribution. In this case, the relevant primal objective function is defined by $h(x)=\1_{\{x:\,e^{\T}x\geq t\}}$, where $t\in\R$ is a fixed level. More generally, $h(x)=\1_{\{x:\,\Psi(x)\geq t\}}$ can be chosen, where $\Psi$ is a relevant risk aggregation function, or $h(x)$ can model any risk measure of choice. Our next example is the Marginal Problem with Copula Bounds, an extension to the marginal problem mentioned in \cite{puccetti1}. The copula defined by the probability measure $\calF$ with marginals $F_i$ is the function \begin{align*} \calC_{\calF}:\,[0,1]^n&\rightarrow[0,1],\\ u&\mapsto F\left(F_1^{-1}(u_1),\dots,F_n^{-1}(u_n)\right). \end{align*} A copula is any function $\calC:[0,1]^n\rightarrow[0,1]$ that satisfies $\calC=\calC_{\calF}$ for some probability measure $\calF$ on $\R^n$. Equivalently, a copula is the multivariate cdf of any probability measure on the unit cube $[0,1]^n$ with uniform marginals. In quantitative risk management, using the model \begin{equation*} \quad\sup_{\calF\in\cM_{F_1,\dots,F_n}}\,\int_{\R^n}h(x)\diff\calF(x) \end{equation*} to bound the worst case risk for a random vector $X$ with marginal distributions $F_i$ can be overly conservative, as no dependence structure between the coordinates of $X_i$ is assumed given at all. The structure that determines this dependence being the copula $\calC_{\calF}$, where $\calF$ is the multivariate distribution of $X$, Embrechts \& Puccetti \cite{puccetti1} suggest problems of the form \begin{align*} \text{(P')}\quad\sup_{\mu\in\cM_{F_1,\dots,F_n}}\,&\int_{\R^n}h(x)\diff\mu(x), \\ \text{s.t. }&\calC_{\text{lo}}\leq\calC_{\calF}\leq\calC_{\text{up}}, \end{align*} as a natural framework to study the situation in which partial dependence information is available. In problem (P'), $\calC_{\text{lo}}$ and $\calC_{\text{up}}$ are given copulas, and inequality between copulas is defined by pointwise inequality, \begin{equation*} \calC_{\text{lo}}(u)\leq\calC_{\calF}(u)\quad(u\in[0,1]^n). \end{equation*} Once again, (P') is a special case of the general framework studied in Section \ref{duality}, as it is equivalent to write \begin{align*} \text{(P)}\quad\sup_{\calF}\,&\int_{\R^n}h(x)\diff\calF(x)\\ \text{s.t. }&\int_{\R^n}\1_{\{x\leq(F_1^{-1}(u_1),\dots,F_n^{-1}(u_n))\}} (u,x)\diff\calF(x)\leq\calC_{\text{up}}(u),\quad(u\in[0,1]^n),\\ &\int_{\R^n}-\1_{\{x\leq(F_1^{-1}(u_1),\dots,F_n^{-1}(u_n))\}} (u,x)\diff\calF(x)\leq-\calC_{\text{lo}}(u),\quad(u\in[0,1]^n),\\ &\int_{\R^n}\1_{\{x_i\leq z\}} (z,x)\diff\calF(x)=F_i(z),\quad(i\in\N_n,\,z\in\R),\\ &\calF\geq 0. \end{align*} The dual of this problem is given by \begin{align*} \text{(D)}\quad\inf_{\calG_{\text{up}},\calG_{\text{lo}},\calS_1,\dots,\calS_n}\,& \int_{[0,1]^n}\calC_{\text{up}}(u)\diff\calG_{\text{up}}(u)- \int_{[0,1]^n}\calC_{\text{lo}}(u)\diff\calG_{\text{lo}}(u) +\sum_{i=1}^n\int_{\R}F_i(z)\diff\calS_i(z)\\ \text{s.t. }&\int_{[0,1]^n}\1_{\{x\leq(F_1^{-1}(u_1),\dots,F_n^{-1}(u_n))\}} (u,x)\diff\calG_{\text{up}}(u)\\ &\hspace{1cm}-\int_{[0,1]^n}\1_{\{x\leq(F_1^{-1}(u_1),\dots,F_n^{-1}(u_n))\}} (u,x)\diff\calG_{\text{lo}}(u)\\ &\hspace{1cm}+\sum_{i=1}^n\int_{\R}\1_{\{x_i\leq z\}}\diff\calS_i(z) \geq h(x),\quad(x\in\R^n),\\ &\calG_{\text{lo}},\calG_{\text{up}}\geq 0. \end{align*} Using the notation $s_i, S_i$ introduced in Section \ref{special cases}, this problem can be written as \begin{align*} \inf_{\calG_{\text{up}},\calG_{\text{lo}},\calS_1,\dots\calS_n}\,& \int_{[0,1]^n}\calC_{\text{up}}(u)\diff\calG_{\text{up}}(u)- \int_{[0,1]^n}\calC_{\text{lo}}(u)\diff\calG_{\text{lo}}(u) +\sum_{i=1}^n\int_{\R}(s_i-S_i(z))\diff F_i(z)\\ \text{s.t. }&\calG_{\text{up}}(\cB(x)) -\calG_{\text{lo}}(\cB(x)) +\sum_{i=1}^n(s_i-S_i(x_i))\geq h(x),\quad(x\in\R^n),\\ &\calG_{\text{up}},\calG_{\text{lo}}\geq 0, \end{align*} where $\cB(x)=\{u\in[0,1]^n:\,u\geq(F_1(x_1),\dots,F_n(x_n))\}$. To the best of our knowledge, this dual has not been identified before. Due to the high dimensionality of the space of variables and constraints both in the primal and dual, the marginal problem with copula bounds is difficult to solve numerically, even for very coarse disrecte approximations. \section{Robust Risk Aggregation via Bounds on Integrals} \label{bounds on integrals} In quantitative risk management, distributions are often estimated within a parametric family from the available data. For example, the tails of marginal distributions may be estimated via extreme value theory, or a Gaussian copula may be fitted to the multivariate distribution of all risks under consideration, to model their dependencies. The choice of a parametric family introduces {\em model uncertainty}, while fitting a distribution from this family via statistical estimation introduces {\em parameter uncertainty}. In both cases, a more robust alternative would be to study models in which the available data is only used to estimate upper and lower bounds on finitely many integrals of the form \begin{equation}\label{of the form} \int_{\Phi}\phi(x)\diff\calF(x), \end{equation} where $\phi(x)$ is a suitable test function. A suitable way of estimating upper and lower bounds on such integrals from sample data $x_i$ $(i\in\N_k)$ is to estimate confidence bounds via bootstrapping. \subsection{Motivation}\label{motivation} To motivate the use of constraints in the form of bounds on integrals \eqref{of the form}, we offer the following explanations: First of all, discretized marginal constraints are of this form with piecewise constant test functions, as the requirement that $F_i(\xi_{k})-F_i(\xi_{k-1})=b_k$ $(k=1,\dots,\ell)$ for a fixed set of discretization points $\xi_0<\dots<\xi_{\ell}$ can be expressed as \begin{equation}\label{new form} \int_{\Phi}1_{\{\xi_{k}\leq x_i\leq\xi_{k-1}\}}\diff\calF(x)=b_k,\quad(k=1,\dots,\ell). \end{equation} It is furthermore quite natural to relax each of these equality constraints to two inequality constraints \begin{equation*} b^{\ell}_{k,i}\leq\int_{\Phi}1_{\{\xi_{k}\leq x_i\leq\xi_{k-1}\}}\diff\calF(x)\leq b^u_{k,i} \end{equation*} when $b_k$ is estimated from data. More generally, constraints of the form $\probability[X\in S_j]\leq b^u_j$ for some measurable $S_j\subseteq\R^n$ of interest can be written as \begin{equation*} \int_{\Phi} 1_{S_j}(x)\diff\calF(x)\leq b^u_j. \end{equation*} A collection of $\ell$ constraints of this form can be relaxed by replacing them by a convex combination \begin{equation*} \int_{\Phi}\sum_{j=1}^{\ell}w_j 1_{S_j}(x)\diff\calF(x)\leq\sum_{j=1}^{\ell} w_j b^u_j, \end{equation*} where the weights $w_j>0$ satisfy $\sum_j w_j =1$ and express the relative importance of each constituent constraint. Non-negative test functions thus have a natural interpretation as importance densities in sums-of-constraints relaxations. This allows one to put higher focus on getting the probability mass right in regions where it particularly matters (e.g., values of $X$ that account for the bulk of the profits of a financial institution), while maximzing the risk in the tails without having to resort to too fine a discretization. While this suggests to use a piecewise approximation of a prior estimate of the density of $X$ as a test function, the results are robust under mis-specification of this prior, for as long as $\phi(x)$ is nonconstant, constraints that involve the integral \eqref{of the form} tend to force the probability weight of $X$ into the regions where the sample points are denser. To illustrate this, consider a univariate random variable with density $f(x)=2/3(1+x)$ on $x\in[0,1]$ and test function $\phi(x)=1+a x$ with $a\in[-1,1]$. Then $\int_{0}^{1}\phi(x) f(x)\diff x=1+5a/9$. The most dispersed probability measure on $[0,1]$ that satisfies \begin{equation}\label{hallo} \int_{0}^{1}\phi(x)\diff\calF(x)=1+\frac{5a}{9} \end{equation} has an atom of weight $4/9$ at $0$ and an atom of weight $5/9$ at $1$ independently of $a$, as long as $a\neq 0$. The constraint \eqref{hallo} thus forces more probability mass into the right half of the interval $[0,1]$, where the unknown (true) density $f(x)$ has more mass and produces more sample points. As a second illustration, take the density $f(x)=3 x^2$ and the same linear test function as above. This time we find $\int_{0}^{1}\phi(x) f(x)\diff x=1+3a/4$, and the most dispersed probability measure on $[0,1]$ that satisfies \begin{equation*} \int_{0}^{1}\phi(x)\diff\calF(x)=1+\frac{3a}{4} \end{equation*} has an atom of weight $3/4$ at $0$ and an atom of weight $1/4$ at $1$ independently of $a\neq 0$, with similar conclusions as above, except that the effect is even stronger, correctly reflecting the qualitative features of the density $f(x)$. \subsection{General Setup and Duality}\label{general setup} Let $\Phi$ be decomposed into a partition $\Phi=\bigcup_{i=1}^k\Xi_i$ of polyhedra $\Xi_i$ with nonempty interior, chosen as regions in which a reasonable number of data points are available to estimate integrals of the form \eqref{of the form}. Each polyhedron has a primal description in terms of generators, \begin{equation*} \Xi_i=\conv(q^i_1,\dots,q^i_{n_i})+\cone(r^i_1,\dots,r^i_{o_i}) \end{equation*} where $\conv(q^i_1,\dots,q^i_{n_i})$ is the polytope with vertices $q^i_n\in\R^n$, and \begin{equation*} \cone(r^i_1,\dots,r^i_{o_i})=\left\{\sum_{m=1}^{o_i}\xi_m r^i_m:\, \xi_m\geq 0\;(m\in\N_{o_i})\right\} \end{equation*} is the polyhedral cone with recession directions $r^i_m\in\R^n$. Each polyhedron also has a dual description in terms of linear inequalities, \begin{equation*} \Xi_i=\bigcap_{j=1}^{k_i}\left\{x\in\R^n:\, \langle f^i_{j}, x\rangle\geq \ell^i_j\right\}, \end{equation*} for some vectors $f^i_j\in\R^n$ and bounds $\ell^i_j\in\R$. The main case of interest is where $\Xi_i$ is either a finite or infinite box in $\R^n$ with faces parallel to the coordinate axes, or an intersection of such a box with a linear half space, in which case it is easy to pass between the primal and dual descriptions. Note however that the dual description is preferrable, as the description of a box in $\R^n$ requires only $2n$ linear inequalities, while the primal description requires $2^n$ extreme vertices. Let us now consider the problem \begin{align*} \text{(P)}\quad\sup_{\calF\in\cM_{\fF}}\,&\int_{\Phi}h(x)\diff \calF(x)\\ \text{s.t. }&\int_{\Phi}\phi_{s}(x)\diff\calF(x) \leq a_{s},\quad(s=1,\dots,M),\\ \text{s.t. }&\int_{\Phi}\psi_{t}(x)\diff\calF(x) = b_{t},\quad(t=1,\dots,N),\\ &\int_{\Phi}1\diff\calF(x)=1,\\ &\calF\geq 0, \end{align*} where the test functions $\psi_t$ are piecewise linear on the partition $\Phi=\bigcup_{i=1}^k\Xi_i$, and where $-h(x)$ and the test functions $\phi_s$ are piecewise linear on the infinite polyhedra of the partition, and either jointly linear, concave, or convex on the finite polyhedra (i.e., polytopes) of the partition. The dual of (P) is \begin{align} \text{(D)}\quad\inf_{(y,z)\in\R^{M+N+1}} \,&\sum_{s=1}^{M}a_s y_s+\sum_{t=1}^{N}b_t z_t + z_0,\nonumber\\ \text{s.t. }&\sum_{s=1}^{M}y_s\phi_s(x)+\sum_{t=1}^{N}z_t\psi_t(x)+z_0\1_{\Phi}(x) -h(x)\geq 0,\; (x\in\Phi),\label{pos0}\\ &y\geq 0.\nonumber \end{align} We remark that (P) is a semi-infinite programming problem with infinitely many variables and finitely many constraints, while (D) is a semi-infinite programming problem with finitely many variables and infinitely many constraints. However, the constraint \eqref{pos0} of (D) can be rewritten as copositivity requirements over the polyhedra $\Xi_i$, \begin{equation*} \sum_{s=1}^{M}y_s\phi_s(x)+\sum_{t=1}^{N}z_t\psi_t(x)+z_0\1_{\Phi}(x) -h(x)\geq 0,\quad(x\in\Xi_i), \quad(i=1,\dots,k). \end{equation*} Next we will see how these copositivity constraints can be handled numerically, often by relaxing all but finitely many constraints. Nesterov's first order method can be adapted to solve the resulting problems, see \cite{nesterov1, nesterov2, sergey}. In what follows, we will use the notation \begin{equation*} \varphi_{y,z}(x)=\sum_{s=1}^{M}y_s\phi_s(x)+\sum_{t=1}^{N}z_t\psi_t(x)+z_0-h(x). \end{equation*} \subsection{Piecewise Linear Test Functions}\label{piecewise linear} The first case we discuss is when $\phi_s|_{\Xi_i}$ and $h|_{\Xi_i}$ are jointly linear. Since we furthermore assumed that the functions $\psi_t|_{\Xi_i}$ are linear, there exist vectors $v^i_s\in\R^n$, $w^i_t\in\R^n$, $g^i\in\R^n$ and constants $c^i_s\in\R$, $d^i_t\in\R$ and $e^i\in\R$ such that \begin{align*} \phi_s|_{\Xi_i}(x)&=\langle v^i_s, x\rangle+c^i_s,\\ \psi_t|_{\Xi_i}(x)&=\langle w^i_t, x\rangle+d^i_t,\\ h|_{\Xi_i}(x)&=\langle g^i, x\rangle+e^i. \end{align*} The copositivity condition \begin{equation*} \sum_{s=1}^{M}y_s\phi_s(x)+\sum_{t=1}^{N}z_t\psi_t(x)+z_0\1_{\Phi}(x) -h(x)\geq 0,\quad(x\in\Xi_i) \end{equation*} can then be written as \begin{multline*} \langle f^i_j, x\rangle \geq \ell^i_j,\quad (j=1,\dots,k_i)\Longrightarrow\\ \left\langle\sum_{s=1}^{M}y_s v^i_s+\sum_{t=1}^{N}z_t w^i_t-g^i\;,\; x\right\rangle\geq e^i-\sum_{s=1}^{M}y_s c^i_s - \sum_{t=1}^{N}z_t d^i_t - z_0. \end{multline*} By Farkas' Lemma, this is equivalent to the constraints \begin{align} \sum_{s=1}^{M}y_s v^i_s+\sum_{t=1}^{N}z_t w^i_t-g^i&=\sum_{j=1}^{k_i}\lambda^i_j f^i_j,\label{c1}\\ e^i-\sum_{s=1}^{M}y_s c^i_s - \sum_{t=1}^{N}z_t d^i_t - z_0&\leq\sum_{j=1}^{k_i}\lambda^i_j \ell^i_j,\label{c2}\\ \lambda^i_j&\geq 0,\quad(j=1,\dots,k_i),\label{c3} \end{align} where $\lambda^i_j$ are additional auxiliary decision variables. Thus, if all test functions are linear on all polyhedral pieces $\Xi_i$, then the dual (D) can be solved as a linear programming problem with $M+N+1+\sum_{i=1}^k k_i$ variables and $k(n+1)$ linear constraints, plus bound constraints on $y$ and the $\lambda^i_j$. More generally, if some but not all polyhedra correspond to jointly linear test function pieces, then jointly linear pieces can be treated as discussed above, while other pieces can be treated as discussed below. Let us briefly comment on numerical implementations, further details of which are described in the second author's thesis \cite{sergey}: An important case of the above described framework corresponds to a discretized marginal problem in which $\phi_s(x)$ are piecewise constant functions chosen as follows for $s=(i,j)$, $(\iota=1,\dots,n; j=1,\dots,m)$: Introduce $m+1$ breakpoints $\xi^{\iota}_0<\xi^{\iota}_1<\dots<\xi^{\iota}_m$ along each coordinate axis $\iota$, and consider the infinite slabs \begin{equation*} S_{\iota,j}=\left\{x\in\R^n:\, \xi^{\iota}_{j-1}\leq x_{\iota}\leq\xi^{\iota}_{j}\right\}, \quad(j=1,\dots,m). \end{equation*} Then choose $\phi_{\iota,j}(x)=1_{S_{\iota,j}}(x)$, the indicator function of slab $S_{\iota,j}$. We remark that this approach corresponds to discretizing the constraints of the Marginal Problem described in Section \ref{special cases}, but not to discretizing the probability measures over which we maximize the aggregated risk. While the number of test functions is $nm$ and thus linear in the problem dimension, the number of polyhedra to consider is exponentially large, as all intersections of the form \begin{equation*} \Xi_{\iota,\vec{j}}=\bigcap_{\iota=1}^n S_{\iota, j_{\iota}} \end{equation*} for the $m^n$ possible choices of $\vec{j}\in\N_{m}^n$ have to be treated separately. In addition, in VaR applications $h(x)$ is taken as the indicator function of an affine half space $\{x:\sum x_\iota \geq \tau\}$ for a suitably chosen threshold $\tau$, and for CVaR applications $h(x)$ is chosen as the piecewise linear function $h(x)=\max(0, \sum x_\iota -\tau)$. Thus, polyhedra $\Xi_{\iota,\vec{j}}$ that meet the affine hyperplane $\{x:\,\sum x_\iota=\tau\}$ are further sliced into two separate polyhedra. A straightforward application of the above described LP framework would thus lead to an LP with exponentially many constraints and variables. Note however, that the constraints \eqref{c1}--\eqref{c3} now read \begin{align} g^i&=\sum_{j=1}^{k_i}\lambda^i_j f^i_j,\label{c4}\\ e^i-\sum_{s=1}^{M}y_s c^i_s - z_0&\leq\sum_{j=1}^{k_i}\lambda^i_j \ell^i_j,\label{c5}\\ \lambda^i_j&\geq 0,\quad(j=1,\dots,k_i),\label{c6} \end{align} as $v^i_s=0$ and no test functions $\psi_t(x)$ were used, with $g^i=[\begin{smallmatrix}1&\dots&1\end{smallmatrix}]^{\T}$ when $\Xi_i\subseteq\{x:\, \sum x_{\iota}\geq\tau\}$ and $g^i=0$ otherwise. That is, the vector that appears in the left-hand side of Constraint \eqref{c4} is fixed by the polyhedron $\Xi_i$ alone and does not depend on the decision variables $y,z_0$. Since $z_0$ is to be chosen as small as possible in an optimal solution of (D), the constraint \eqref{c5} has to be made as slack as possible. Therefore, the optimal values of $\lambda^i_j$ are also fixed by the polyhedron $\Xi_i$ alone and are identifiable by solving the small-scale LP \begin{align*} (\lambda^{i}_j)^*=\arg\max_{\lambda}\;&\sum_{j=1}^{k_i}\lambda^{i}_j\ell^{i}_j\\ \text{s.t. }-g^i&=\sum_{j=1}^{k_i}\lambda^i_j f^i_j,\\ \lambda^i_j&\geq 0,\quad(j=1,\dots,k_i). \end{align*} In other words, when the polyhedron $\Xi_i$ is considered for the first time, the variables $(\lambda^i_j)^*$ can be determined once and for all, after which the constraints \eqref{c4} -- \eqref{c6} can be replaced by \begin{equation*} e^i-\sum_s y_{s}c^i_{s}-z_0\leq C_i, \end{equation*} where $C_i=\sum_{j=1}^{k_i}(\lambda^i_j)^* \ell^i_j$ and where the sum on the left-hand side only extends over the $n$ indices $s$ that correspond to test functions that are non-zero on $\Xi_i$. Thus, only the $nm+1$ decision variables $(y,z_0)$ are needed to solve (D). Furthermore, the exponentially many constraints correspond to an extremely sparse constraint matrix, making the dual of (D) an ideal candidate to apply the simplex algorithm with delayed column generation. A similar approach is possible for the situation where $\phi_s$ is of the form \begin{equation*} \phi_{s}(x)=1_{S_{s}}(x)\times\left(\langle v_{s}, x\rangle + c_{s}\right), \end{equation*} for all $s=(\iota,j)$. The advantage of using test functions of this form is that fewer breakpoints $\xi_{\iota,j}$ are needed to constrain the distribution appropriately. To illustrate the power of delayed column generation, we present the following numerical experiments that compare the computation times of our standard simplex (SS) implementation and our simplex with delayed column generation (DCG) for different dimensions $d$ and numbers $k$ of polyhedra : \begin{table}[h!] \begin{center} \begin{tabular}{|c|c|c|c|c|} \hline $d$ & $k$ & DCG & SS & DCG/SS\\ \hline $2$ & $256$ & $2.282$ & $0.963$ & $2.340$\\ $2$ & $1'024$ & $1.077$ & $1.345$ & $0.801$\\ $2$ & $4'096$ & $3.963$ & $7.923$ & $0.500$ \\ $2$ & $16'384$ & $20.237$ & $34.495$ & $0.587$ \\ $2$ & $65'536$ & $124.89$ & $187.826$ & $0.665$ \\ \hline $3$ & $4'096$ & $2.653$ & $13.103$ & $0.202$ \\ $3$ & $32'768$ & $42.863$ & $145.002$ & $0.300$ \\ \hline $4$ & $4'096$ & $4.792$ & $25.342$ & $0.189$ \\ $4$ & $65'536$ & $345.6$ & $28'189.7$ & $0.012$ \\ \hline \end{tabular} \end{center} \end{table} The results show that with increasing dimension, the savings of using delayed column generation over the standard simplex algorithm grows rapidly. Since our Matlab implementation of both methods are not optimized, the absolute computation times can be considerably improved, but the speedup achieved by using delayed column generation is representative. \subsection{Piecewise Convex Test Functions}\label{piecewise convex} When $\phi_s|_{\Xi_i}$ and $-h|_{\Xi_i}$ are jointly convex, then $\varphi_{y,z}(x)$ is convex. The copositivity constraint \begin{equation*} \sum_{s=1}^{M}y_s\phi_s(x)+\sum_{t=1}^{N}z_t\psi_t(x)+z_0\1_{\Phi}(x) -h(x)\geq 0,\quad(x\in\Xi_i) \end{equation*} can then be written as \begin{equation*} \langle f^i_j, x\rangle \geq \ell^i_j,\quad (j=1,\dots,k_i)\Longrightarrow \varphi_{y,z}(x)\geq 0, \end{equation*} and by Farkas' Theorem (see e.g., \cite{polik}), this condition is equivalent to \begin{align} \varphi_{y,z}(x)+\sum_{j=1}^{k_i}\lambda^i_j\left(\ell^i_j-\langle f^i_j, x\rangle\right)&\geq 0,\quad(x\in\R^n),\label{global}\\ \lambda^i_j&\geq 0,\quad(j=1,\dots,k_i),\nonumber \end{align} where $\lambda^i_j$ are once again auxiliary decision variables. While \eqref{global} does not reduce to finitely many constraints, the validity of this condition can be checked numerically by globally minimizing the convex function $\varphi_{y,z}(x)+\sum_{j=1}^{k_i}\lambda^i_j\left(\ell^i_j-\langle f^i_j, x\rangle\right)$. The constraint \eqref{global} can then be enforced explicitly if a line-search method is used to solve the dual (D). \subsection{Piecewise Concave Test Functions}\label{piecewise concave} When $\phi_s|_{\Xi_i}$ and $-h|_{\Xi_i}$ are jointly concave but not linear, then $\varphi_{y,z}(x)$ is concave and $\Xi_i=\conv(q^i_1,\dots,q^i_{n_i})$ is a polytope. The copositivity constraint \begin{equation}\label{copos} \sum_{s=1}^{M}y_s\phi_s(x)+\sum_{t=1}^{N}z_t\psi_t(x)+z_0\1_{\Phi}(x) -h(x)\geq 0,\quad(x\in\Xi_i) \end{equation} can then be written as \begin{equation*} \varphi_{y,z}(q^i_j)\geq 0,\quad(j=1,\dots,n_i). \end{equation*} Thus, \eqref{copos} can be replaced by $n_i$ linear inequality constraints on the decision variables $y_s$ and $z_t$. \subsection{Piecewise Polynomial Test Functions}\label{piecewise polynomial} Another case that can be treated via finitely many constraints is when $\phi_s|_{\Xi_i}$, $\psi_t|_{\Xi_i}$ and $h|_{\Xi_i}$ are jointly polynomial. The approach of Lasserre \cite{lasserre} and Parrilo \cite{parrilo} can be applied to turn the copositivity constraint \begin{equation*} \langle f^i_j, x\rangle \geq \ell^i_j,\quad (j=1,\dots,k_i)\Longrightarrow \varphi_{y,z}(x)\geq 0, \end{equation*} into finitely many linear matrix inequalities. However, this approach is generally limited to low dimensional applications. \\ \section{Conclusions} Our analysis shows that a wide range of duality relations in use in quantitative risk management can be understood from the single perspective of a generalized duality relation discussed in Section \ref{duality}. An interesting class of special cases is provided by formulating a finite number of constraints in the form of bounds on integrals. The duals of such models are semi-inifinite optimization problems that can often be reformulated as finite optimization problems, by making use of standard results on co-positivity. \\ \noindent {\bf Acknowledgements:} Part of this research was conducted while the first author was visiting the FIM at ETH Zurich during sabbatical leave from Oxford. He thanks for the support and the wonderful research environment he encountered there. The research of the the first two authors was supported through grant EP/H02686X/1 from the Engineering and Physical Sciences Research Council of the UK.
\section{Introduction} \label{Sect_1} Eruptive young stellar objects (YSOs) are characterized by dramatically increased accretion from the circumstellar disk onto the star \citep{HK96}. In addition to the classical FU Orionis \citep[FUor;][]{Herbig77,HK96,Reipurth10} and EX Lupi \citep[EXor;][]{Herbig07,Herbig08,Lorenzetti12} type stars, observations of the past decades revealed several embedded eruptive young stars whose classification is uncertain, and differ from both classical types in amplitude and time scale \citep{Reipurth10,Kun11a,Kospal13}. Although the outbursts in each class are powered by enhanced accretion from the disk onto the star, the spectroscopic properties of FUors and EXors are radically different. EXors keep the major spectroscopic signatures of magnetospheric accretion during the outburst, whereas FUor spectra indicate star--disk interaction zones substantially transformed by the outburst. The powerful winds, associated with the high accretion rate, produce Herbig--Haro (HH) objects, which may thus be important tracers of the accretion history \citep{Reipurth10}. Since a sizeable part of the stellar mass may build up during repeated outbursts \citep{VB06}, and the high accretion luminosity and associated outflows affect the structure and evolution of the protoplanetary disks, outbursting stars are key objects for understanding the formation of Sun-like stars and their planetary systems. Their extreme rarity and diversity, however, hinder the understanding of the origin and nature of the outbursts. The recent review by \citet{Audard14} lists 26 FUors and FUor-like objects, and 18 EXors can be found in the list of \citet{Lorenzetti12}. New discoveries and their subsequent monitoring are thus relevant to a better understanding of the phenomenon of episodic accretion. The new candidate eruptive stars presented in this Letter are found in Lynds~1340, an isolated dark cloud at ({\it l,b\/}) = (130.1\degr,11.5\degr). The first large-scale studies of the region \citep{KOS94,KAY03,MMN03} suggested that L1340, located at 600\,pc from the Sun, has formed a few mid-B, A and early F type, and some two dozens of low-mass stars. We conducted a comprehensive multi-wavelength study of L1340, involving {\it Spitzer\/}, {\it WISE\/}, {\it 2MASS\/}, and {\it SDSS\/} photometric data, a search for \ha\ emission stars via slitless grism spectroscopy, long-slit optical spectroscopy, narrow-band optical and high angular resolution near infrared imaging observations (M. Kun et al. 2014, in preparation). This work led to a revised distance of $730\pm30$\,pc for L1340, and resulted in the discovery of some 250 candidate YSOs. Among them we found three new candidate eruptive young stars, which deserve special attention. One of them is IRAS~02224+7227, identified by \citet{KAY03} as the probable exciting source of HH~487, situated at 6.2\arcmin\ southwest of the star. The two others are 2MASS~02263797+7304575 and 2MASS~02325605+7246055, neither of them was mentioned before as a young star. We present the observational results which prove the eruptive nature of these objects. We adopt the revised distance of 730\,pc. \section{Data} \label{Sect_data} Lynds~1340 was observed by the \textit{Spitzer Space Telescope} using the Infrared Array Camera \citep[IRAC;][]{Fazio2004} on 2009 March 16 and the Multiband Imaging Photometer for Spitzer \citep[MIPS;][]{Rieke2004} on 2008 November 26 (Prog. ID: 50691, P.I. G. Fazio). The observations covered $\sim 1$~deg$^2$ in each band. The centers of the 3.6 and 5.8\,\mum\ images are slightly displaced from those of the 4.5 and 8\,\mum\ images, therefore the south-eastern part of the molecular cloud was not covered by the 4.5 and 8\,\mum\ maps. Moreover, the south-western part of L1340 is outside the field of view of the 24 and 70\,\mum\ images. We performed aperture photometry on individual saturation corrected basic calibrated data (CBCD) IRAC images, produced by the S18.18.0 pipeline at the Spitzer Science Center (SSC), using a 2-pixel aperture radius, and a sky annulus between 2 and 6 pixels. An additional array-dependent photometric correction and a pixel-phase correction \citep[see][]{Hora2008}, as well as an aperture correction were applied. The final photometry and its uncertainty were estimated as the average and rms of the individual flux densities measured in different CBCD frames. For MIPS data we used MOPEX \citep{Makovoz2006} to create mosaic maps from the 24\,\mum\ enhanced BCD and 70\,\mum\ BCD products (version S18.13.0) of SSC. Following \citet{Gordon2007} at 70\,\mum\ before mosaicing we made a column mean subtraction and a time filtering on the BCD frames. PSF photometry was used to determine the targets' fluxes in the mosaic maps. The resulted fluxes of the three targets are shown in Table\,\ref{Tab1}. The quoted uncertainties were computed as a quadratic sum of the measurement errors and absolute calibration errors of 2\% for IRAC \citep{Hora2008} and 4 and 7\% for 24 and 70\,\mum\ MIPS observations \citep{Engelbracht2007,Gordon2007}, respectively. We obtained low resolution optical spectra of the star coinciding with IRAS~02224+7227 on 2003 February 5 using CAFOS\footnote{\url{http://w3.caha.es/CAHA/Instruments/CAFOS/}} with the G--100 grism on the 2.2-m telescope of the Calar Alto Observatory, and on 2004 December 11 using FAST on the 1.5-m FLWO telescope \citep{Fabricant}. We reduced and analysed the spectra in IRAF. High angular resolution {\it JHK\/} images, centered on the same star, were obtained on 2002 October 24 using the near-infrared camera Omega-Cass\footnote{\url{http://www.caha.es/CAHA/Instruments/OCASS/ocass.html}}, mounted on the 3.5-m telescope of the Calar Alto Observatory. The plate scale was 0.1\arcsec/pixel. The target was observed at four dithering positions, and the observations consisted of two dither cycles. The total integration time was 480\,s in each filter. We reduced and analysed the data in IRAF. Following the flat-field correction and bad pixel removal the sky frame for each cycle was obtained by taking the minimum of the images at different dithering positions. This sky frame was subtracted from each individual image of a given cycle, and the frames from a single cycle were combined into a mosaic image. Aperture photometry was performed on the reduced images. The instrumental magnitudes were transformed into the {\it JHK}$_\mathrm{s}$ system by using the \tm\ magnitudes of field stars within the field of view. The resulting magnitudes are listed in Table~\ref{Tab1}. We performed a new search for \ha\ emission stars in L1340 using the Wide Field Grism Spectrograph~2 ({\it WFGS2}) installed on the University of Hawaii 2.2-meter telescope. The instrument setup and data reduction procedure were same as described in detail in \citet{Szeg13}. We observed 2MASS~02263797+7304575 on 2011 October 16, and detected a \ha\ emission with $EW(\ha)=-80$~\AA\ in its spectrum. The \ks\ magnitude of 2MASS~02325605+7246055 was measured on the images obtained on 2010 October 18, during the monitoring program of V1180~Cas \citep{Kun11a}, using the MAGIC camera on the 2.2-m telescope of the Calar Alto Observatory. Narrow band images through $[$\ion{S}{2}$]$ and \ha\ filters, as well as broad \textit{R}-band images containing the environment of 2MASS~02325605+7246055, were obtained with the Schmidt telescope of the Th\"uringer Landessternwarte (TLS), Tautenburg in May, June and September 2011. The frames obtained through the same filter were coadded, leading to integration times of 0.2, 3.0, and 4.0 hours in the $R$, \ha, and $[$\ion{S}{2}$]$, respectively. Spectra of the nebula and the two brightest HH knots were obtained using the TLS medium-resolution Nasmyth spectrograph ($R\sim700$) in 2011 November. The total integration time was 2 hours per object. {\it BVR$_\mathrm{C}$I$_\mathrm{C}$\/} photometric observations of IRAS~02224+7227 were performed with the 1-m Ritchey--Chretien--Coude (RCC) telescope of the Konkoly Observatory at three epochs between 2001 and 2011. We measured the {\it R}$_\mathrm{C}$ and {\it I}$_\mathrm{C}$ magnitudes of IRAS~02224+7227 and 2MASS~02263797+7304575 at several epochs between 2011 January and 2014 June on the images collected with the wide-field camera on the Schmidt telescope of the Konkoly Observatory to monitor the light variations of V1180~Cas \citep{Kun11a}. The results of the photometry are listed in the machine-readable version of Table~\ref{Tab1}. L1340 is situated within Stripe~1260 of the {\it SEGUE\/} survey \citep{Yanny2009}, thus its whole area was observed in the {\it ugriz\/} bands in 2005 November--December. Each target star has high-quality 3.4, 4.6, 12, and 22 micron fluxes in the \wise\ data base. These data offer useful pieces of information on their spectral energy distribution (SED) and long-term photometric behavior. \section{Results and Discussion} \subsection{IRAS~02224+7227: a FUor-like Star} IRAS~02224+7227 (2MASS~02270555+7241167, HH\,487\,S) is situated near the south-western edge of L1340. The upper panel of Fig.~\ref{Fig1} shows its FAST and CAFOS spectra, together with that of FU~Ori, found in the FAST Public Archive\footnote{\url{http://tdc-www.harvard.edu/cgi-bin/arc/fsearch}}, and with the spectrum of the G5-type supergiant star HD~47731, found in \citeauthor{LeBorgne}'s \citeyearpar{LeBorgne} spectrum library. The hydrogen and metallic absorption spectra, characteristic of G-type stars, the \ion{Sr}{2} line at 4077\,\AA, indicative of high luminosity class, and the youth-indicator \ion{Li}{1} line at 6707\,\AA\ can clearly be identified, suggesting the FU~Ori nature of this star. Comparison of the strength of the H$\gamma$, H$\beta$ lines and the G-band with those in spectroscopic standards \citep{JHC84,LeBorgne} suggests G4--G5 spectral type. Weak \ha\ emission can be seen in the FAST spectrum, while the \ha\ line is in absorption in the CAFOS spectrum. We estimated a foreground extinction of $A_\mathrm{V} \approx 2.0$~mag from the G5 spectral type, and the ({\it R\/}$_\mathrm{C}-${\it I\/}$_\mathrm{C}$ ) color index, adopting an intrinsic color index of ({\it R\/}$_\mathrm{C}-${\it I\/}$_\mathrm{C})_0= 0.37$ \citep{HK96}. The second panel of Fig.~\ref{Fig1} shows the SED of the star, constructed from \sdss\ {\it u\/}, \citep[transformed into Johnson {\it U\/} following][]{Jordi06}, \spitzer, \tm, \wise, and our own {\it BVR$_\mathrm{C}$I$_\mathrm{C}$\/} data. IRAS~02224+7227 lies outside the field of view of the 70\,\mum\ MIPS image, and only flux upper limits are available for the IRAS 60 and 100\,\mum\ bands. The shape of the dereddened SED in the 0.36--3.6\,\mum\ region can be satisfactorily approximated with the sum of a G5I and an M6 type photosphere, contributing nearly equally to the total flux at 2\,\mum, and supporting that FUors exhibit wavelength-dependent spectral types \citep{HK96}. Integrating the dereddened SED over the 0.36--24\,\mum\ wavelength interval, and assuming a distance of 730~pc we obtain a luminosity of $L(0.36-24) \approx 23$\,$L_{\sun}$. Including the {\it IRAS\/} 60\,\mum\ and 100\,\mum\ flux upper limits we obtain $L(0.36-100) < 59$\,$L_{\sun}$. Although most of the known FUors have $L_\mathrm{bol} > 100 L_{\sun}$, both values are within the range of FUor luminosities ($7 \la L_\mathrm{bol}/L_{\sun} \la 800$) compiled by \citet{Audard14}. The third panel of Fig.~\ref{Fig1} shows the central part of the Omega-Cass {\it K\/}-band image. IRAS~02224+7227 has three wide companions. Table~\ref{Tab2} lists the separations from the central star, position angles (from the North towards the East), and {\it JHK$_\mathrm{s}$} magnitudes of the three stars. Each companion has normal stellar colors in the near-infrared bands. The angular separations correspond to 1750--4000~AU at 730\,pc, comparable with the typical size of protostellar envelopes. Spectroscopy and/or L/M-band photometry are required to decide whether these stars are physically related to each other or not. Since the brightness of the star in the POSS\,1 image, recorded on 1954 September 27, is similar to the more recent ones shown in Table~\ref{Tab1}, IRAS~02224+7227 has been in outburst for at least sixty years. Similarly to other FUor-like stars whose outburst dates are unknown, we observe in this star an evolved phase of outburst. If HH~487 was created by a previous outburst of the star and we assume a typical HH-object space velocity 200~km\,s$^{-1}$, then its angular distance of 6.2\arcmin, corresponding to 1.3\,pc without accounting for the unknown inclination, suggests that a previous outburst might have happened some 6500 years ago. This interval corresponds to statistical estimates on the average time span between outbursts \citep{Scholz13}. The present outburst thus might have occurred in a disk shaped by previous outburst(s). The Class~II SED, absence of bright reflection nebulosity, and the relatively low luminosity may also suggest an `old' FUor. Our photometric measurements indicate no trend in the optical magnitudes between 2001 and 2014 (Table~\ref{Tab1}). \subsection{2MASS 02263797+7304575: a possible EXor} We identified this star as a candidate classical T~Tauri star (CTTS) based on the strong \ha\ emission ($EW(\ha)=-80$~\AA) in its spectrum. The shape of its SED, constructed from the {\it NOMAD\/} {\it BVRI\/}, \sdss, \tm, \wise, and \spitzer\ data, and plotted in Fig.~\ref{Fig2} (left), together with the band of typical SEDs of the Taurus pre-main sequence stars \citep{DAlessio}, confirms the CTTS nature. The \sdss\ {\it griz\/} magnitudes, measured on 2005 November 3, were transformed into the {\it BVR$_\mathrm{C}$I$_\mathrm{C}$\/} system \citep{Ivezic07} to compare them with other available photometric data. It is apparent that while the optical magnitudes of the {\it NOMAD\/} catalogue, the \tm, \spitzer, and \wise\ data, measured at various epochs, smoothly delineate a Class~II SED, the \sdss\ magnitudes stand apart, suggesting that the star was unusually bright over the optical spectrum on 2005 November 3. We estimated the spectral type, extinction, and luminosity of the star by comparing the short-wavelength (\textit{BVRIJ}) side of the low-state SED with a grid of reddened photospheres, using the pre-main sequence colors and bolometric corrections tabulated by \citet{Pecaut2013} and the extinction law of \citet{CCM89}, as well as the $A_\mathrm{V} \ge 0.5$\,mag \citep[foreground extinction toward L1340,][]{KWT03} restriction. The dashed line shows the best estimate, the SED of a K4-type photosphere, fitted to the {\it NOMAD\/} and \tm~{\it J\/} magnitudes, dereddened by $A_\mathrm{V}=0.7$\,mag. Its photospheric luminosity is 0.28\,$L_{\sun}$ at a distance of 730\,pc. The total luminosity of the system, derived by integrating the dereddened SED and extrapolating the contribution of the spectral regions beyond 70\,\mum\ using the method by \citet{Chavarria}, is $L_\mathrm{bol} \approx 0.39$\,$L_{\sun}$, suggesting $L_\mathrm{disk} / L_\mathrm{star} \approx 0.39$, higher than the upper limit ($\sim0.2$) for passive irradiated disks \citep{KH87}. The {\it R}$_\mathrm{C}$ vs. {\it R}$_\mathrm{C}-${\it I}$_\mathrm{C}$ color--magnitude diagram is plotted in Fig.~\ref{Fig2} (right). It can be seen that, with the exception of the \sdss\ point, the star's {\it R}$_\mathrm{C}$ magnitudes stay in the $15.4 <${\it R}$_\mathrm{C} < 16.3$~mag interval. The amplitude of the light variations within an observing season is about 0.5\,mag, and the average brightness changes from season to season. The distribution of the points suggests variable circumstellar extinction. The single bright point indicates a burst-like event with an amplitude of 1.5--2.0~mag. Similar outbursts are supposed to frequently occur in CTTSs, but due to their short duration are hard to catch. Detection of such events may be helpful in looking for the specialities of EXor disks, distinguishing them from those of less violent pre-main sequence stars. Such specialities may be the high $L_\mathrm{disk} / L_\mathrm{star}$, and the flattening of the SED between 24 and 70\,\mum, unlike typical T~Tauri SEDs, and indicative of a remnant infalling envelope \citep{Calvet94}. A similar flattening can be seen in the SED of prototype EXor, EX Lupi \citep{Sipos}, and the candidate EXors identified by \citet{Giannini} also exhibit flat SEDs near the ClassI/ClassII boundary, indicating that eruptive events occur at the earliest evolutionary phases of the protoplanetary disks. \subsection{2MASS~02325605+7246055: an Outbursting Protostellar Object?} This star is invisible in the DSS\,1 red image. A faint bow-shaped nebula appears near its position in the DSS\,2 red image, indicating that the outburst of an embedded star opened a cavity in its circumstellar envelope between 1954 September 27 and 1994 January 3. The nebula is also visible in our {\it R}$_\mathrm{C}$ and {\it I}$_\mathrm{C}$ images, obtained since 2001 with various instruments of the Tautenburg and Konkoly Observatories. The shortest wavelength where the star appears is the K-band. Our \ks\ image obtained in 2010, as well as the \spitzer\ 3.6 and 4.5\,\mum\ images (Fig.~\ref{Fig3}, upper right panel) show a small fan-shaped nebula next to the star at the northwestern side, similar to those found near several embedded eruptive stars (e.g. RNO~125 associated with PV~Cep \citep{Kun11b}, and McNeil's Nebula with V1647~Ori \citep{Briceno04}). The \ks\ magnitude of the star, measured in 2010, was some 0.5 mag brighter than the \tm\ \ks\ (Table~\ref{Tab1}), suggesting a years time-scale brightening of the star. Our \ha\ and $[$\ion{S}{2}$]$ images (Fig.~\ref{Fig3}, upper left) show two faint HH knots, located at 71\arcsec\ (SE) and 112\arcsec\ (NW) to the northwest of the star, at (RA, Dec)(SE)=($2^\mathrm{h}32^\mathrm{m}45.3^\mathrm{s}; +72\degr46\arcmin56\arcsec$), and (RA, Dec)(NW)=($2^\mathrm{h}32^\mathrm{m}40.2^\mathrm{s}; +72\degr47\arcmin32\arcsec$), respectively. The HH knots are associated with shocked H$_2$ emission seen in \textit{IRAC\/} 4.5-micron image. While the spectrum of the \textit{SE} HH knot has low signal-to-noise ratio, the \textit{NW} knot exhibits \ha, $[$\ion{O}{1}$]$, and $[$\ion{S}{2}$]$ emission lines (Fig.~\ref{Fig3}, lower left panel). The radial velocity of the \ha\ is $v_\mathrm{LSR} = -85\pm30$\,km\,s$^{-1}$, consistent with the cometary morphology of the object in the sense that we see the scattered light from the blueshifted outflow lobe. The monopolar morphology in the optical/IR implies a significant inclination. The spectrum of the object shows a faint continuum but lacks \ha\ emission. This indicates that at present accretion is weak or even absent which points to the lack/replenishment of the inner disk, most probably depleted during the outburst. The SED, plotted in the lower right panel of Fig.~\ref{Fig3}, suggests an embedded object with deep silicate absorption at 10\,\mum. Its bolometric luminosity, estimated from the 70\,\mum\ flux following the method of \citet{Dunham}, is $L_\mathrm{bol} \approx 1.2$\,$L_{\sun}$ at 730\,pc. The lack of \ha\ emission from the spectrum and the low bolometric luminosity suggest the present low-accretion state of 2MASS~02325605+7246055. The remnant of its recent outburst, the reflection nebula which appeared between 1954 and 1995, suggests an EXor-like event. \section{Conclusions} We identified three candidate eruptive young stars associated with the molecular cloud L1340, whose YSO population consists of some 250 members (Kun et al. 2014, in prep.). Together with the previously identified V1180~Cas \citep{Kun11a}, some 1.6\% of the total YSO poulation belong to eruptive classes. The observed properties of the FUor-like IRAS~02224+7227 suggest a late stage of its episodically accreting phase of evolution. The EXor candidate 2MASS~02263757+7304575 exhibits a Class~II SED, but its high 70\,\mum\ flux is indicative of a remnant infalling envelope. The nebulosity beside the Class~I protostar 2MASS~02325605+7246055 is a signature of a recent EXor-like outburst. These results indicate that amplitudes and time scales of YSO outbursts do not strictly correlate with the evolutionary stage of the star. The divergent observed properties of the eruptive young stars \citep[e.g.][]{Quanz07,Lorenzetti12} point to objects at various episodes of their consecutive outbursts. \begin{acknowledgements} We are grateful to G\'abor F\H{u}r\'esz for obtaining the FAST spectrum of IRAS~02224+7227. This work makes use of observations made with the \textit{Spitzer Space Telescope}, which is operated by the Jet Propulsion Laboratory, California Institute of Technology under a contract with NASA. Our results are partly based on observations collected at the Centro Astron\'omico Hispano Alem\'an (CAHA) at Calar Alto, operated jointly by the Max-Planck Institut f\"ur Astronomie and the Instituto de Astrof\'{\i}sica de Andaluc\'{\i}a (CSIC). This research utilized observations with the $2.2$-m telescope of the University of Hawaii and we thank Colin Aspin and Bo reipurth for their interest and support. Our research has benefited from the VizieR catalogue access tool, CDS, Strasbourg, France. Financial support from the Hungarian OTKA grant K81966 is acknowledged. \end{acknowledgements}
\subsection*{Acknowledgements} \noindent The author would like to thank to A. Dedes, M. Paraskevas and K. Suxho for collaboration in the calculations of the rare top quark decay rates. This work was supported in part by the Polish National Science Center under the research grants DEC-2011/01/M/ST2/02466 and DEC-2012/05/B/ST2/02597.
\section{Introduction} \label{intro} Massive vector particles are of relevance for various subfields of physics including the weak interaction, test models for photon-mass searches, and the regularization of certain infrared divergences in quantum-field calculations. In this latter context, the mass term should be introduced such that the symmetries of the original model remain unspoiled. For a U(1) Lorentz-invariant gauge theory, this can be accomplished with the Stueckelberg method.\cite{Stueckelberg} In recent years, Lorentz-violating quantum field theories have become a focus of theoretical\cite{SMEtheo} and experimental\cite{SMEexp} inquiry. Such theories also require the regularization of infrared divergences in some circumstances, and the question arises, as to whether the Stueckelberg method can be adapted to such Lorentz-breaking U(1) gauge theories. As in the Lorentz-invariant case, the adapted Stueckelberg method should be compatible with all symmetries of the theory to be regulated, so that the usual Stueckelberg approach can be employed for the Lorentz-violating extension of conventional QED. However, the Lorentz violation in this model allows a broader range of compatible Stueckelberg terms that may be introduced, and this freedom can then be used to streamline calculations in Lorentz-breaking QED. The present work provides one possible class of extensions of the usual Stueckelberg procedure to Lorentz-violating QED. Section \ref{lagr} gives a brief overview of the procedure at the lagrangian level. Some implications of the Lorentz-breaking Stueckelberg Lagrangian are discussed in Sec.~\ref{EoM}. \section{Lagrangian Analysis} \label{lagr} We begin with the usual free-photon Lagrangian in the minimal Standard-Model Extension coupled to a conserved source $j^\mu$: \begin{eqnarray} \label{lagrangian} {\cal L}_{\gamma} \!& = &\! -\tfrac{1}{4}F^2 -A\cdot j -\tfrac{1}{4}(k_F)^{\kappa\lambda\mu\nu} F_{\kappa\lambda}F_{\mu\nu} +\tfrac{1}{2}\epsilon^{\kappa\lambda\mu\nu}(k_{AF})_\kappa A_\lambda F_{\mu\nu}\,. \end{eqnarray} Here, Lorentz and CPT breakdown is controlled by the spacetime-constant backgrounds $k_F$ and $k_{AF}$. The absence of Lorentz symmetry in the above Lagrangian~(\ref{lagrangian}) allows us to drop the requirement of Lorentz invariance for the mass-type term $\delta {\cal L}_{m}$ to be introduced for the photon. However, the inclusion of arbitrary Lorentz violation into $\delta {\cal L}_{m}$ may be problematic: consider the case in which $k_F$ and $k_{AF}$ are such that a subgroup of the Lorentz group remains unbroken. A regulator violating this residual symmetry may be undesirable, so that the breakdown of the remaining invariant subgroup may have to be excluded from $\delta {\cal L}_{m}$. For our present purposes, however, we consider arbitrary Lorentz violation in $\delta {\cal L}_{m}$. This yields more general results also relevant for purposes other than infrared regularization. More specifically, we implement the Stueckelberg procedure by introducing a scalar field $\phi$ as follows:\cite{Cam12} \begin{equation} \label{Stueckelberg} \delta {\cal L}_{m}= \tfrac{1}{2}(\partial_\mu \phi-m A_\mu) \hat{\eta}^{\mu\nu}(\partial_\nu \phi-m A_\nu)\,, \end{equation} where $m$ can later be identified with the photon mass and \begin{equation} \label{hat-eta} \hat{\eta}^{\mu\nu}=\eta^{\mu\nu}+\hat{G}^{\mu\nu} \end{equation} with $\hat{G}^{\mu\nu}$ a Lorentz-violating operator that may contain derivatives but is otherwise spacetime constant. The small number of Lorentz-symmetric pieces still contained in an arbitrary $\hat{G}^{\mu\nu}$ can be removed if necessary. As in the conventional Stueckelberg case, the key feature of ${\cal L}_\gamma+\delta {\cal L}_{m}$ is its invariance (up to total derivatives) under a local gauge transformation \begin{equation} \label{eq:gaugetransf} \delta A_\mu = \partial_\mu \epsilon(x)\,,\quad \delta \phi = m \epsilon(x)\,. \end{equation} With the addition of $\xi$-type gauge fixing $\mathcal{L}_\textrm{g.f.}=-\frac{1}{2\xi} (\partial_\mu \hat{\eta}^{\mu \nu}A_\nu + \xi m \phi)^2$ and a Faddeev--Popov contribution $\mathcal{L}_\textrm{F.P.}$, the model Lagragian ${\cal L}={\cal L}_\gamma+\delta {\cal L}_{m}+\mathcal{L}_\textrm{g.f.}+ \mathcal{L}_\textrm{F.P.}$ becomes\cite{Cam12} \begin{eqnarray} \label{eq:totalL} {}\hspace{-10mm}\mathcal{L}\! &=&\!-\tfrac{1}{4}F^2 -A\cdot j +\tfrac{1}{2}m^2 A_\mu \hat{\eta}^{\mu \nu} A_\nu -\frac{1}{2\xi}(\partial_\mu \hat{\eta}^{\mu \nu}A_\nu)^2\nonumber\\ &&\! {}-\tfrac{1}{2}\phi(\partial_\mu\hat{\eta}^{\mu\nu}\partial_\nu+\xi m^2)\phi - \bar c(\partial_\mu\hat{\eta}^{\mu\nu}\partial_\nu+\xi m^2)c\frac{{}}{{}}\nonumber\\ &&\!{}-\tfrac{1}{4}(k_F)^{\kappa\lambda\mu\nu} F_{\kappa\lambda}F_{\mu\nu} +\tfrac{1}{2}\epsilon^{\kappa\lambda\mu\nu}(k_{AF})_\kappa A_\lambda F_{\mu\nu}\,. \end{eqnarray} Note that the scalar $\phi$ and the ghosts $c$ and $\bar c$ are now decoupled and can be integrated out. We can therefore disregard these fields in what follows. \section{Equations of Motion} \label{EoM} The Lagrangian~(\ref{eq:totalL}) yields the following equation of motion for the photon: \begin{eqnarray} \label{EoM1} \left[\eta^{\mu\alpha}\eta^{\nu\beta}\partial_\mu +(k_{AF})_\mu\epsilon^{\mu\nu\alpha\beta} +(k_F)^{\mu\nu\alpha\beta}\partial_\mu\right]F_{\alpha\beta}&&\nonumber\\ {}+\left[m^2\hat{\eta}^{\mu\nu} +\frac{1}{\xi}\hat{\eta}^{\mu\alpha}\hat{\eta}^{\nu\beta}\partial_\alpha\partial_\beta\right]A_\mu &=& j^\nu \,. \end{eqnarray} Taking the 4-divergence of this equation yields \begin{equation} \label{longitudinal} \left(\hat{\eta}^{\mu\nu}\partial_\mu\partial_\nu+\xi m^2 \right)(\partial_\alpha\hat{\eta}^{\alpha\beta}A_\beta)=0\,. \end{equation} Note that $(\partial_\alpha\hat{\eta}^{\alpha\beta}A_\beta)$ projects out one degree of freedom contained in $A_\mu$. Equation~(\ref{longitudinal}) establishes that this degree of freedom is not excited by the source $j_\nu$, so $(\partial_\alpha\hat{\eta}^{\alpha\beta}A_\beta)$ is an auxiliary mode. This is consistent with the expectation of three physical degrees of freedom for a massive vector field. A plane-wave ansatz in Eq.~(\ref{EoM1}) yields the model's dispersion relation; it has the following structure:\cite{Cam12} \begin{equation} \label{fullDR} \frac{1}{\xi} \big(1+\tfrac{1}{4}\hat{G}^\alpha_\alpha\big) (\hat{\eta}^{\mu\nu}p_\mu p_\nu-\xi m^2)Q(p)=0\,. \end{equation} The first factor does not contain physical modes. The $(\hat{\eta}^{\mu\nu}p_\mu p_\nu-\xi m^2)$ piece corresponds to the auxiliary mode governed by Eq.~(\ref{longitudinal}). The factor $Q(p)$ is associated with the physical degrees of freedom and has the structure \begin{equation} \label{physDR} Q=(p^2-m^2)^3+ r_2(p^2-m^2)^2+ r_1(p^2-m^2)+ r_0\,. \end{equation} Here, the coefficients $r_j$ are coordinate scalars containing the Lorentz-breaking tensors $k_F$ and $k_{AF}$. They vanish in the limit $k_F, k_{AF}\to0$. The explicit expressions for the $r_j$ can be found in the literature.\cite{Cam12} Note that the physical dispersion relation~(\ref{physDR}) is consistent with the expectation of three conventional massive modes perturbed by small Lorentz violation. The exact expression for the corresponding propagator is somewhat unwieldy.\cite{Cam12} But for most applications only leading-order Lorentz-violating effects need to be taken into account. The Lorentz-breaking contributions can then be incorporated via this propagator insertion (see also Fig.~\ref{fig}):\cite{Cam12} \begin{eqnarray} \label{delta-S} \delta S(p)^{\mu\nu} & = & 2i(k_{AF})_\alpha\epsilon^{\alpha\beta\mu\nu}p_\beta -2(k_{F})^{\alpha\mu\beta\nu}p_\alpha p_\beta\nonumber\\ &&{}+m^2\hat{G}^{\mu\nu} -\frac{1}{\xi}\big(\hat{G}^\mu_\alpha\, p^\alpha p^{\nu}+\hat{G}^\nu_\alpha\, p^\alpha p^{\mu}\big)\,. \end{eqnarray} \begin{figure}[t] \begin{center} \psfig{file=insertion.eps,width=2.0in} \end{center} \caption{Propagator insertion. The wavy lines denotes the conventional Lorentz-invariant Stueckelberg propagator $P_0$. The square box represents the leading-order Lorentz-breaking insertion given by Eq.~(\ref{delta-S}). } \label{fig} \end{figure} \section*{Acknowledgments} This work has been supported in part by the Indiana University Center for Spacetime Symmetries, by Universidad Andr\'es Bello under Grant No.~UNAB DI-27-11/R, as well as by the Portuguese Funda\c c\~ao para a Ci\^encia e a Tecnologia.
\section{Introduction and Preliminaries} Totally reflexive modules were introduced by Auslander and Bridger \cite {AusBr} as modules of Gorenstein dimension zero. It was not until 2002 when Avramov and Martsinkovsky \cite{AvMar} first referred to them as totally reflexive, to better emphasize their homological properties. Over Gorenstein rings, totally reflexive modules are exactly the maximal Cohen-Macaulay modules, whose representation theory is well developed, for example see \cite{LWbook}, \cite{Yosbook}. However, over non-Gorenstein rings much less is known. Something that is known from \cite{sing} is that if there exists a non-free totally reflexive module over a non-Gorenstein local ring with the residue field of characteristic 0, then there exist infinitely many non-isomorphic indecomposable totally reflexive modules over the ring as well. In fact in \cite {BT} and \cite {Holm} infinite families of non-isomorphic indecomposable totally reflexive modules are constructed, both of which arise from exact zero divisors. Although totally reflexive modules were originally defined for a broader class of rings, for this paper we assume $R$ to be a commutative Noetherian local ring. A finitely generated $R$-module $M$ is called \emph{totally reflexive} if the biduality map, $\delta: M \rightarrow \operatorname{Hom}_R(\operatorname{Hom}_R(M,R), R)$ is an isomorphism, $\operatorname{Ext}_R^i(M,R) =0 $ and $\operatorname{Ext}_R^i(\operatorname{Hom}_R(M,R), R) =0 $ for all $i>0.$ Projective modules are obviously totally reflexive. We call a nonzero totally reflexive module \emph{nontrivial} if it is not a projective module. A complex is called \emph {acyclic} if its homology is zero. A \emph{totally acyclic complex} is an acyclic complex $\mathbb{A}$ whose \emph{dual} $\operatorname{Hom}_R(\mathbb{A}, R)$ is also acyclic. A module being totally reflexive is equivalent to being a syzygy in a totally acyclic complex of finitely generated free modules. Recall that over a local ring $(R, \mathfrak{m})$ a complex $(\mathbb{F}, \partial)$ of free modules is said to be \emph{minimal} if $\operatorname{im} \partial_i^{\mathbb{F}} \subseteq \mathfrak{m}\mathbb{F}_{i-1}$ for all $i.$ Through out this paper we will investigate these modules via their presentation matrix, which is a matrix whose columns are a minimal generating set of the module. Presentation matrices are not unique. However, when over commutative Noetherian local rings two minimal presentation matrices of the same module must be equivalent, see \cite{equivmat}. In this paper, we investigate the structure of totally reflexive modules over local non-Gorenstein Artinian rings by looking at their totally reflexive submodules. We are most interested in the case when the quotient module formed by the totally reflexive module and a totally reflexive submodule is also totally reflexive and of minimal length. If this occurs, then we say that the module has a \emph{saturated TR-filtration,} see Definition \ref{TRfiltDefn} for a more precise description. We prove that knowing if a totally reflexive module has a saturated TR-filtration is directly linked to the existence of a minimal presentation matrix of the module that is upper triangular. \\ \begin{textbf}{Theorem}\end{textbf}. \emph{Let $(R,\mathfrak{m})$ be a non-Gorenstein ring with $\mathfrak{m}^3=0\neq \mathfrak{m}^2 $ that contains exact zero divisors, and suppose that $T$ is a totally reflexive $R$-module. There exists a saturated TR-filtration of $T$ if and only if $T$ has an upper triangular presentation matrix.\\} From this theorem, we obtain two corollaries. Corollary \ref{cor1} gives further information about the form of this upper triangular presentation matrix. Additionally, Corollary \ref{UTres} proves the existence of a complete resolution in which every differential can be simultaneously represented by upper triangular matrices. We conclude with an extensive example of the theory, including a study of the number of ways a saturated TR-filtration can occur. \subsection{Previously Known Results} Since the results in this paper are proven with the ring being commutative local non-Gorenstein with the cube of the maximal ideal equaling zero, we give two previously known theorems which provide useful facts about these types of rings and totally reflexive modules over them. For a local Artinian ring $(R, \mathfrak{m}, k),$ set $e=\operatorname{length} (\mathfrak{m}/\mathfrak{m}^2)$ to be the embedding dimension of $R.$ \begin{theorem} \cite[Theorem 3.1]{Yos} \label{YosThm} Let $(R,\mathfrak{m},k)$ be a commutative local Artinian non-Gorenstein ring with $\mathfrak{m}^3=0\neq\mathfrak{m}^2.$ If there exist a nontrivial totally reflexive $R$-module $T,$ then the following conditions hold: \begin{enumerate}[(a)] \item $(0 :_R \mathfrak{m})=\mathfrak{m}^2$ \item $\operatorname{length} (0 :_R \mathfrak{m})=e-1,$ in particular, $\operatorname{length}(R)=2e.$ \label{YosThmlength} \item Let $n$ be the minimal number of generators of $T,$ then $\operatorname{length} (T)=ne.$ \item $T$ has a minimal free resolution of the form \label{YosThmReso} $$\cdots \rightarrow R^n \xrightarrow{d_2} R^n \xrightarrow{d_1}R^n \rightarrow T \rightarrow 0.$$ \end{enumerate} \end{theorem} \begin{remark} Part (\ref{YosThmReso}) of the previous theorem implies that every differential in a minimal free resolution of $T$ can be presented by a square matrix. Moreover, any minimal presentation matrix of $T$ is a square matrix. This fact holds for a complete resolution of $T$ as well, since any syzygy in a complete resolution of a totally reflexive modules is also totally reflexive. \end{remark} \begin{theorem}\cite[Theorem 5.3]{BT} \label{BTcite} Let $(R, \mathfrak{m})$ be a local ring with $\mathfrak{m}^3 = 0$ and $e \geq 3.$ Let $x$ be an element of $\mathfrak{m}/\mathfrak{m}^2;$ the following conditions are equivalent. \begin{enumerate} [(i)] \item The element $x$ is an exact zero divisor in $R$. \item The Hilbert series of $R$ is $1+e\tau +(e-1)\tau^2,$ and there exists an exact sequence of finitely generated free R-modules $$ \mathbb{F}: \qquad F_3\rightarrow F_2 \xrightarrow{} F_1 \xrightarrow{\psi}F_0 \rightarrow F_{-1} $$ such that $\operatorname{Hom}_R(\mathbb{F},R)$ is exact, the homomorphisms are represented by matrices with entries in $\mathfrak{m},$ and $\psi$ is represented by a matrix in which some row has $x$ as an entry and no other entry from $\mathfrak{m}/\mathfrak{m}^2.$ \end{enumerate} \end{theorem} An exact zero divisor is a special type of ring element, one that is quite significant in these results, see Definition \ref {defnexpair}. \section {Filtrations and Upper Triangular Presentation Matrices} Before we present the main theorem of this paper, we need to make note of a few facts when considering modules via their presentation matrices. \begin{lemma} \label{copyk} Let $(R,\mathfrak{m},k)$ be a commutative local non-Gorenstein ring with $\mathfrak{m}^3=0\neq\mathfrak{m}^2$ and $M$ an $R$-module. If a presentation matrix of $M$ has a column whose entries are contained in $\mathfrak{m}^2,$ then the first syzygy of $M$ contains a copy of $k$ as a direct summand. \end{lemma} \begin{proof} Let $\mathfrak{m}=(x_1, \dots , x_e),$ where $e$ is the embedding dimension of $R,$ and suppose $M$ has the presentation matrix $\mathbf{M}= [c_1, \dots, c_s],$ where $c_i \in R^{b_0}.$ Also, after row and column operations, we may assume that $c_1 \subset \mathfrak{m}^2 R^{b_0}$. Let $$\mathbb{F}: \qquad \cdots \rightarrow R^{b_2}\xrightarrow{\mathbf{N}} R^{b_1}\xrightarrow{\mathbf{M}} R^{b_0}\rightarrow 0$$ be a deleted free resolution of $M$. Since $\mathbf{MN}=0,$ for $1\leq i\leq e$ we have that the elements $\left(\begin{array}{c} x_i\\ 0\\ \vdots \\ 0\end{array} \right),$ are part of a minimal generating set of the syzygies of $\mathbf{M}.$ Without loss of generality, assume that $\mathbf{N}$ has the form $$\left[\begin{array}{cccccc} x_1 &\ldots & x_e& 0& \ldots &0\\ 0 &\ldots & 0 & *& \ldots &*\\ \vdots& \ddots& \vdots&\vdots& \ddots& \vdots\\ 0 &\ldots& 0 & *& \ldots &*\\ \end{array} \right]. $$ Note that the entries in the first row, after the $e$th column, are all zeroes. This is true since if there was a nonzero entry there, then it would be a linear combination of the first $e$ columns. Therefore, $\operatorname{coker} \mathbf{N} \cong k \oplus X,$ for some $R$-module $X$. \end{proof} \begin{proposition}\label{ksumcor} Let $(R,\mathfrak{m},k)$ be a non-Gorenstein ring with $\mathfrak{m}^3=0\neq\mathfrak{m}^2.$ If the $n$th syzygy of an $R$-module $M$, $\Omega^R_n(M)$ contains a copy of $k$ as a direct sum, then $M$ is not totally reflexive. \end{proposition} \begin{proof} Assume that for some $R$-module $X,$ we have $\Omega^R_n(M)\cong k \oplus X$. Now suppose that $M$ is totally reflexive, and therefore, $\Omega^R_n(M)$ is totally reflexive as well since it is a syzygy of $M$. Then, for all $i>0,$ we have $\operatorname{Ext}^i_R(\Omega^R_n(M),R)=0 $ and thus $\operatorname{Ext}^i_R(k\oplus X,R)=0$ for all $i>0,$ which implies that $\operatorname{Ext}^i_R(k,R)=0,$ for all $i>0$. This holds if and only if $R$ is Gorenstein. However, we assumed $R$ not to be Gorenstein, and therefore $M$ cannot be totally reflexive. \end{proof} \begin{definition}\label{TRfiltDefn} For a totally reflexive $R$-module $T$, a \emph{TR-filtration} of $T$ is a chain of submodules $$0 = T_0 \subset T_1 \subset \dots \subset T_{n-1} \subset T_n = T,$$ in which the following hold for all $i=1, \dots, n:$ \begin{enumerate}[(i)] \item $T_i$ is totally reflexive \item $T_{i}/T_{i-1}$ is totally reflexive \item $T_{i}/T_{i-1}$ contains no proper nonzero totally reflexive submodules.\\ \noindent We say a filtration is a \emph{saturated TR-filtration} if, in addition, we have that \item $T_{i}/T_{i-1}$ is of minimal length among the totally reflexive $R$-modules. \end{enumerate} \end{definition} In \cite{HenSega} the elements of the ring which play a vital role in the construction of these filtrations are defined as follows. \begin{definition}\label{defnexpair} For a commutative ring $A$, a non-unit $a \in A$ is said to be an \emph {exact zero divisor} if there exists $b\in A$ such that $(0: a)= (b)$ and $(0: b)= (a).$ If $A$ is local then $b$ is unique up to unit, and we call $(a,b)$ an \emph {exact pair of zero divisors}.\\ This is equivalent to the existence of a free resolution of $A/(a)$ of the form $$ \cdots \rightarrow A \xrightarrow{[b]}A \xrightarrow{[a]}A \xrightarrow{[b]}A \xrightarrow{[a]}A \rightarrow 0$$ \begin{remark} \label{TAexact} The complex in Definition \ref{defnexpair} is totally acyclic and thus the modules $A/(a)$ and $A/(b)$ are totally reflexive. In fact, in certain cases the reverse it also true. \end{remark} \end{definition} \begin{lemma} \label{expairisTR} Let $(R, \mathfrak{m})$ be a local ring with $\mathfrak{m}^3=0 \neq\mathfrak{m}^2,$ and let $a\in R.$ If $R/(a)$ is a totally reflexive module, then $a$ is an exact zero divisor. \end{lemma} \begin{proof} Let $R/(a)$ be a totally reflexive module, and so by Theorem \ref{BTthm} it has a free resolution of the form $$\mathbb{F}: \qquad \cdots \rightarrow R \xrightarrow{b_2}R \xrightarrow{b_1}R \xrightarrow{a} R \rightarrow 0.$$ This implies that $(0:a)=(b_1).$ Apply the $\operatorname{Hom}$ functor, we obtain the complex $$\operatorname{Hom}_R(\mathbb{F},R): 0 \rightarrow \operatorname{Hom}_R(R/(a),R) \xrightarrow{}R \xrightarrow{a^*}R \xrightarrow{b^*_1} R \rightarrow \cdots,$$ which is exact because $R/(a)$ is totally reflexive. This is isomorphic to $$ 0 \rightarrow (R/(a))^* \xrightarrow{}R \xrightarrow{a}R \xrightarrow{b_1} R \rightarrow \cdots.$$ Therefore, $\ker(a)=(b_1)$ and we have that $(0:b_1)=(a).$ Hence $(a,b_1)$ are an exact pair of zero divisors. \end{proof} \begin{lemma}\label {lem1} Let $(R,\mathfrak{m})$ be a non-Gorenstein ring with $\mathfrak{m}^3=0\neq\mathfrak{m}^2$ and embedding dimension $e.$ If a totally reflexive $R$-module $T$ has a minimal presentation matrix that contains a row with only one nonzero entry, then there exists a totally reflexive submodule $U \subset T$ such that length $(U) = $ length $(T) -e.$ \\ \end{lemma} \begin{proof} Let $$\mathbb{F}: \qquad \cdots \rightarrow F_2\xrightarrow{\mathbf{W}} F_1 \xrightarrow{\mathbf{T}} F_0 \rightarrow F_{-1} \rightarrow \cdots$$ be a complete free resolution of $T,$ where $\mathbf{W}=(\omega_{ij})$ is a presentation matrix of $\Omega^R_1(T)$ and let $$\mathbf{T}=\left[ \begin{array}{cccc} t_{11} &\cdots & & t_{1n} \\ \vdots & & &\vdots\\ t_{n-1 1} &\cdots & & t_{n-1 n}\\ 0& \cdots &0 & t_{nn}\\ \end{array} \right] $$ \noindent be a presentation matrix of $T.$ From Theorem \ref{BTcite} we know that $t_{nn}$ is an exact zero divisor of $R.$ Let $s$ generate $(0:t_{nn}).$ Since $\mathbf{TW}=0,$ we have that $t_{nn}\omega_{n j}=0 $ for all $j=1, \dots , n$. Thus, the ideal $(\omega_{n1}, \dots , \omega_{nn}) \subset (0: t_{nn})=(s)$. This, with Lemma \ref{copyk}, implies that for some $i$ we have $\omega_{ni} = s,$ up to units. Therefore, every entry in the $n$th row of $\mathbf{W}$ is a multiple of $s$. We can apply column operations to $\mathbf{W}$ to assume that $$ \mathbf{W} = \left[ \begin{array}{cccc} w_{11} &\cdots && w_{1n} \\ \vdots&&& \vdots\\ 0& \cdots & 0 & s \end{array} \right]$$ is another presentation matrix of $\Omega^R_1(T)$. Define $\textbf{W'} $ to be the $(n-1) $ by $(n-1)$ matrix obtained by deleting the $n$th row and $n$th column from $\textbf{W}. $ Similarly, we define $ \textbf{T'} $ to be the $(n-1) $ by $(n-1)$ matrix obtained by deleting the $n$th row and $n$th column from $\textbf{T}. $ Note that $\mathbf{W'T'}=0,$ and now consider the following commutative diagram: \[ \xymatrixrowsep{1.7pc}\xymatrixcolsep{2.1pc}\xymatrix{ & 0 \ar@{->}[d]^{\qquad}& & 0 \ar@{->}[d]^{\qquad}&& 0 \ar@{->}[d] ^{\qquad}&\\ \mathbb{G}':& R^{n-1 } \ar@{->}[rr]^{ \mathbf{W'}} & & R^{n-1} \ar@{->}[rr]^{ \mathbf{T'}} & & R^{n-1} \ar@{->}[r] &0\\ \\ \mathbb{G}:&R^n \ar@{->}[rr]^{\mathbf{W}} \ar@{<-}[uu]^q && R^n \ar@{->}[rr]^{\mathbf{T}} \ar@{<-}[uu]^q && R^n \ar@{->}[r] \ar@{<-}[uu]^q &0\\ \\ \mathbb{G}'':&R \ar@{->}[rr]^{[w]} \ar@{<-}[uu]^p && R \ar@{->}[rr]^{[t_{nn}]} \ar@{<-}[uu]^p && R \ar@{->}[r] \ar@{<-}[uu]^p&0\\ & 0 \ar@{<-}[u] & & 0 \ar@{<-}[u] && 0 \ar@{<-}[u] &\\ &&&&&&\\ } \] \noindent where $q:=\left[ \begin{array}{ccc} 1& & 0\\ &\ddots & \\ 0& &1\\ 0& \dots &0 \end{array} \right]$ and $p:= \left[\begin{array}{cccc} 0 & \ldots &0 & 1\end {array}\right].$ Thus we have an exact sequence of complexes $$ 0 \rightarrow \mathbb{G} ' \rightarrow \mathbb{G} \rightarrow \mathbb{G} '' \rightarrow 0$$ which yields the following long exact sequence of homology. $$ \cdots \rightarrow \operatorname{H}_1(\mathbb{G} '') \rightarrow \operatorname{H}_0(\mathbb{G} ') \rightarrow \operatorname{H}_0(\mathbb{G}) \rightarrow \operatorname{H}_0(\mathbb{G} '') \rightarrow 0.$$ \noindent Note that $\operatorname{H}_1(\mathbb{G} '') =0$ since $ R\stackrel{\left[ w \right]}{\longrightarrow} R \stackrel{\left[t_{nn}\right]}{\longrightarrow} R$ is exact. Let $\operatorname{coker} \mathbf{T'} $ and so $ U \subset T$. Therefore, $\operatorname{H}_0(\mathbb{G} '')=R/ (t_{nn}) \cong T/U$ and we have the short exact sequence $$ 0 \rightarrow U \rightarrow T \rightarrow T/U \rightarrow 0. $$ To see that $U$ is in fact totally reflexive, note that $T/U \cong R/(t_{nn})$ is totally reflexive since $(t_{nn})$ is an exact zero divisor. This, along with the fact that $T$ is totally reflexive, implies that $U$ is as well, see \cite[Corollary 4.3.5]{GorDim}. From Yoshino's theorem, here listed as Theorem \ref{YosThm}, part (\ref{YosThmlength}), we know that the length of $R/ (t_{nn})$ is the number of its minimal generators times the embedding dimension. Therefore, $\operatorname{length}(R/ (t_{nn}))=e$ and thus $\operatorname{length} (U)=\operatorname{length} (T) -e.$ \end{proof} \begin{theorem}\label {thm1} Let $(R,\mathfrak{m})$ be a non-Gorenstein ring with $\mathfrak{m}^3=0\neq \mathfrak{m}^2 $ that contains exact zero divisors, and suppose that $T$ is a totally reflexive $R$-module. There exists a saturated TR-filtration of $T$ if and only if $T$ has an upper triangular presentation matrix. \end{theorem} \begin{proof} Let $T$ be totally reflexive $R$-module with minimal number of generators $\mu(T)=n$ and suppose there exists a saturated TR-filtering of $T$. To show that $T$ has an upper triangular presentation matrix we will use induction on $\mu(T)$. If $\mu(T)=1,$ then any presentation matrix of $T$ would be of size $ 1 \times 1$ and thus trivially upper triangular. Assuming true for $\mu(T)=n,$ consider the case $\mu(T)=n+1$. By Lemma \ref{lem1} there exists a totally reflexive $R$-module $M \subset T$ such that the sequence $$ 0 \rightarrow M \rightarrow T \rightarrow T/M \rightarrow 0$$ is exact and $T/M$ is cyclic. Define $N:= T/M$ and let $\mathbb{F} $ and $\mathbb{G}$ be free resolutions of $M$ and $N$ respectively. Since $R$ is local with $\mathfrak{m}^3=0,$ we can assume that $F_i=R^n$ and $G_i=R,$ for all $i\geq 0$, \cite{Acyclic}. By applying the Horseshoe Lemma, we get the following diagram: \begin{equation}\label{preset} \xymatrixrowsep{3pc}\xymatrixcolsep{3pc}\xymatrix{ 0\ar@{->}[r] &R^n \ar@{->}[r] \ar@{->}[d]^{d_1^M}& R^n\oplus R \ar@{->}[r]\ar@ {->}[d]^{d_1^T} & R\ar@ {->}[r]\ar@ {->}[d]^{d_1^{N}}\ar@ {->}[dl]^{\sigma_1}& 0\\ 0\ar@{->}[r] &R^n \ar@{->}[r] \ar@{->}[d]^{\epsilon_M}& R^n\oplus R \ar@{->}[r]\ar@ {->}[d]^{\epsilon_T} & R\ar@ {->}[r]\ar@ {->}[d]^{\epsilon_N}\ar@ {->}[dl]^{\sigma_0}& 0\\ 0 \ar@{->}[r] & M \ar@{->}[r] & T \ar@{->}[r]& N \ar@{->}[r]&0\\ &0\ar@{<-}[u]&0\ar@{<-}[u]&0\ar@{<-}[u]& } \end{equation} \noindent where $d_1^T (f,g) \mapsto d_1^M(f) + \sigma_1(g)$. Since $d_1^M(f) \in R^n$ and $\sigma_1 (g) \in R^n\oplus R,$ when we consider them as columns in a presentation matrix of $T$ we have the matrix \begin{equation}\label{prest} \left[\begin{array}{cc} d_1^M& f'\\ 0 & g' \end{array} \right], \end{equation} where $\sigma_1(g)=f'+g'$ for some $f' \in R^n$ and $g' \in R,$ we see that it is upper triangular. This matrix is a presentation matrix of $T$. By induction, the matrix representing $d_1^M(f)$ can be taken to be upper triangular and therefore (\ref{prest}) is upper triangular. \\ Now suppose $T$ has an upper triangular presentation matrix, say $$\left[ \begin{array}{ccc} t_{11} &\cdots & t_{1n} \\ & \ddots &\vdots\\ 0& & t_{nn} \end{array} \right] .$$ Again we will use induction on the minimal number of generators. If $n=2,$ then by Lemma \ref{lem1} there exists a totally reflexive $R$-module $T_1\cong R/(t_{11})$ which is a submodule of $T.$ Thus, $T$ has a saturated TR-filtration $$0=T_0 \subset T_1 \subset T_{2}=T$$ with $T_2/T_1\cong R/(t_{22}).$ Now assume that an $n$-generated totally reflexive module with an upper triangular presentation matrix has a saturated TR-filtration $$0=T_0 \subset T_1 \subset \dots. \subset T_{n-1} \subset T_{n}$$ with $T_i/T_{i-1} $ being one generated and totally reflexive for $i=1,\dots, n.$ Consider an $(n+1)$-generated totally reflexive module $T,$ which has with an upper triangular presentation matrix. By Lemma \ref{lem1}, there exists a totally reflexive module $T_{n}$ such that $T_n$ is a submodule of $T$ and has presentation matrix of the form $$\left[ \begin{array}{ccc} t_{11} &\cdots & t_{1 n} \\ & \ddots &\vdots\\ 0& & t_{n n} \end{array} \right]$$ with $T/T_n \cong R/t_{n+1\,n+1}.$ This, combined with the induction hypothesis, shows that $T$ has a saturated TR-filtration. \end{proof} \begin{definition} An \emph{upper triangular complex} is a complex in which every differential can simultaneously be represented by a square matrix $(a_{ij})$ such that $a_{ij}=0$ when $i>j.$ \end{definition} \begin{corollary}\label{UTres} Let $(R,\mathfrak{m})$ be a non-Gorenstein ring with $\mathfrak{m}^3=0\neq \mathfrak{m}^2 $ that contains exact zero divisors. If $T$ is a totally reflexive $R$-module that has an upper triangular minimal presentation matrix, then it has an upper triangular minimal complete resolution. \end{corollary} \begin{proof} This will be done by induction on the number of minimal generators. Let $T$ be a totally reflexive $R$-module that has an upper triangular minimal presentation matrix with $\mu(T)=n.$ and thus the matrix is of size $n \times n.$ Let $\mathbb{F}$ be a free resolution of $T.$ By \cite[Theorem B]{Acyclic}, we have that $\operatorname{rank}_R F_1=n$ for all $i \in \mathbb{Z}.$ If $\mu(T)=1,$ then the free resolution of $T$ is trivially upper triangular. \\ Assume for $\mu(T)=n$ that $\mathbb{F} $ is an upper triangular minimal free resolution and consider the case when $T$ has a presentation matrix of the form $$\left[ \begin{array}{ccc} t_{11} &\cdots & t_{1 \,n+1} \\ & \ddots &\vdots\\ 0& & t_{n+1 \, n+1} \end{array} \right]. $$ By Theorem \ref{thm1}, there exists an saturated TR-filtration and thus the short exact sequence $$ 0 \rightarrow T_{n} \rightarrow T \rightarrow R/(t_{n+1\, n+1})\rightarrow 0.$$ From the Horseshoe Lemma, a diagram similar to (\ref{preset}) can be obtained and extended to include the first syzygies. By a similar argument to the proof of Theorem \ref{thm1}, we see that the first syzygy of $T$ has an upper triangular presentation matrix. Finally, any syzygy in a free resolution of totally reflexive module is also totally reflexive \cite{GorDim}. Therefore, we can apply this corollary to the $n$th syzygy to see that the $n+1$ the syzygy has an upper triangular presentation matrix. \end{proof} \begin{corollary} \label{cor1} Let $(R,\mathfrak{m})$ be a non-Gorenstein ring with $\mathfrak{m}^3=0\neq \mathfrak{m}^2 $ that contains exact zero divisors. For an $R$-module $M$ that has an upper triangular presentation matrix, the non-zero entries on the main diagonal of that presentation matrix are exact zero divisors if and only if $M$ is totally reflexive. \end{corollary} \begin{proof} Both directions of this proof rely on the existence of a saturated TR-filtration. First, suppose $M$ is totally reflexive and has an upper triangular presentation matrix. Hence, we have a saturated TR-filtration $$0 = M_0 \subset M_1 \subset \dots \subset M_{n-1} \subset M_n = M,$$ as well as short exact sequences $$0\rightarrow M_{i-1} \rightarrow M_i\rightarrow M_{i}/M_{i-1}\rightarrow 0$$ for all $i=2, \cdots, n.$ If $i=2,$ then for some $u, v, \alpha \in R$ we have $$0\rightarrow R/(u) \rightarrow \operatorname{coker} \left[\begin{array}{cc} u& \alpha \\ 0&v \end {array}\right] \rightarrow R/(v)\rightarrow 0.$$ From Lemma \ref{expairisTR} both $u$ and $v$ are exact zero divisors. Assume this holds for $n-1,$ and consider $$0\rightarrow M_{n-1} \rightarrow M_n\rightarrow M_{n}/M_{n-1}\rightarrow 0.$$ By the induction hypothesis, $M_{n-1}$ has a upper triangular presentation matrix with exact zero divisors on the main diagonal. If $\mathbf{M}=(a_{ij})$ is a presentation matrix of $M_n,$ then $M_{n}/M_{n-1}\cong R/(a_{nn}).$ Since $R/(a_{nn})$ is totally reflexive, $a_{nn}$ is an exact zero divisor. Therefore, for all $i,$ $a_{ii}$ are exact zero divisors. Now suppose $M$ has an upper triangular presentation matrix where the non-zero entries on the main diagonal are exact zero divisors. By the inductive part of the proof of Theorem \ref{thm1}, $M$ is totally reflexive. \end{proof} \section{Filtrations and Yoneda $\operatorname{Ext}$} Recall that the Yoneda definition \cite[Appendix A3]{Eisenbud} of $\operatorname{Ext}_R^1(N,M)$ for two $R$-modules $M, N$ gives a correspondence between the elements of $\operatorname{Ext}_R^1(N,M)$ and equivalence classes of exact sequences of the form $$ 0 \rightarrow M \rightarrow X \rightarrow N \rightarrow 0.$$ If an $R$-module $T$ has a TR-filtration, $$0 = T_0 \subset T_1 \subset \dots \subset T_{n-1} \subset T_n = T,$$ then there exist short exact sequences $$0\rightarrow T_{i-1} \rightarrow T_i \rightarrow T_i/T_{i-1} \rightarrow 0$$ and hence an element in $\operatorname{Ext}_R^1(T_i/T_{i-1},T_{i-1}).$ In this section, we investigate the number of possible nonequivalent short exact sequences of the above form for a specific ring. We start with brief discussion on the Yoneda definition of $\operatorname{Ext}^1.$ Let $$\cdots \xrightarrow{} F_2 \xrightarrow{\partial_2}F_1\xrightarrow{\partial_1} F_0\xrightarrow{} N \xrightarrow{} 0$$ be a free resolution of $N,$ and so we have an exact sequence $$ 0 \rightarrow \operatorname{im}\partial_1 \rightarrow F_0 \rightarrow N \rightarrow 0.$$ If we have some $(a +\operatorname{im}\partial^*_1 )\in \operatorname{Ext}_R^1(N,M),$ then $a \in \operatorname{Hom}(F_0, M).$ Define $a '$ to be $a$ restricted to $\operatorname{im} \partial_1$ and $\iota$ to be the natural inclusion from $\operatorname{im} \partial_1$ to $F_0.$ We can obtain the following commutative diagram using the pushout of $a'$ and $\iota.$ \[ \xymatrixrowsep{2.7pc}\xymatrixcolsep{1.8pc}\xymatrix{ 0 \ar@{->}[r]^{\qquad} &\operatorname{im} \partial_1 \ar@{->}[d]^{a'} \ar@{->}[r]^{\quad\iota\quad} &F_0 \ar@{->}[r]\ar@{->}[d] & N \ar@{=}[d] \ar@{->}[r]^{\qquad}& 0 \\ 0 \ar@{->}[r] & M\ar@{->}[r]^{\qquad} & (M\oplus F_0)/I \ar@{->}[r] & N \ar@{->}[r]& 0 \\ }\] \noindent where $I=(-a'(r), \iota(r) | r \in \operatorname{im} \partial_1).$ When considering filtrations, $(M \oplus R^{b_0})/I$ is the next larger submodule on the chain of submodules. In order to gain a better understanding of how this translates to a presentation matrix, we will compute this for the second module in a filtration over a specific ring. For the reminder of this paper, let $S=k[X, Y, Z]/(X^2, Y^2, Z^2, YZ),$ and define $x, y, $ and $z$ to be the image of $X, Y, $ and $Z,$ respectively. Also, let $T_1=S/(x+by+cz)$ and $N=S/(x+dy+fz)$ where $b, c, d, f \in k,$ so they are the cokernal of an exact zero divisor. When considering $\operatorname{Ext}_S^1(N, T_{1})$ by the Yoneda definition, a non-zero element of $\operatorname{Ext}_S^1(N, T_{1})$ corresponded to a short exact sequence of the form $$0 \rightarrow T_1 \rightarrow T_2 \rightarrow N \rightarrow 0.$$ If $\operatorname{Ext}_S^1(N, T_{1})\neq 0,$ then this will allow us to have a saturated TR-filtration of $T_2,$ $$0 \subset T_1 \subset T_{2}.$$ We now will find a presentation matrix for possible $T_2.$ Let $$\mathbb{F}_{N}: \qquad \cdots \xrightarrow{} S \xrightarrow{\partial_3} S\xrightarrow{\partial_2} S \xrightarrow{\partial_1} S \rightarrow 0$$ be a (deleted) minimal free resolution of $N$ where $$\partial_i = \left\{\begin{array} {l l} [x+dy+fz], & \quad \mbox{ if } i \mbox{ is odd} \\ \mbox{$[x-dy-fz],$} & \quad \mbox{ if } i \mbox{ is even. }\\ \end{array} \right.$$ For $(a +\operatorname{im}\partial^*_1) \in \operatorname{Ext}_S^1(N, T_1)$ where $\alpha \in \ker(\operatorname{Hom}_S(\partial_2,T_1)),$ we have that $\alpha \in \operatorname{Hom}_S(S, T_1) \cong T_1.$ We will identify elements of $\operatorname{Hom}_S(S, T_1)$ with $T_1$ under the natural isomorphism. Using the previous description, \[ \xymatrixrowsep{2.7pc}\xymatrixcolsep{1.8pc}\xymatrix{ 0 \ar@{->}[r]^{\qquad} &\operatorname{im} \partial_1 \ar@{->}[d]^{\alpha'} \ar@{->}[r]^{\quad\iota\quad} &S\ar@{->}[r]\ar@{->}[d] & N \ar@{=}[d] \ar@{->}[r]^{\qquad}& 0 \\ 0 \ar@{->}[r] & T_1\ar@{->}[r]^{\qquad} & (T_1\oplus S)/I \ar@{->}[r] & N\ar@{->}[r]& 0 \\ }\] we find that the push out is $T_2:=(T_1 \oplus S)/I.$ To find a presentation matrix for $T_2,$ we need to find a matrix $\mathbf{T_2}$ in which the following complex is exact $$ S^2 \xrightarrow{ \mathbf{T_2} }S^2 \xrightarrow{p} T_2=(T_1 \oplus S)/I \rightarrow 0.$$ Take $e_1, e_2$ to be the standard basis in $S^2$ and define $p(e_1)= (\overline{1}, 0) + I$ and $p(e_2)= (0, 1) + I$. To find the image of $\mathbf{T_2}$ we just need to find the kernel of $p$ which is shown below. $$(x+by+cz) p(e_1) \equiv 0 \qquad \mbox{ and } \qquad -\alpha p(e_1) + (x+dy+fz) p(e_2) \equiv 0$$ where $\alpha \in S/(x+by+cz)$. Therefore, a presentation matrix for $T_2$ is $$\mathbf{T_2}=\left[\begin{array}{c c} x+by+cz & -\tilde{\alpha}\\ 0 & x+dy+fz \end {array} \right],$$ where $\tilde{\alpha}$ is a preimage of $\alpha$ in $S.$ \subsection{The Rank of $\operatorname{Ext}_S^1(N,T_1)$ } Since every element in $\operatorname{Ext}^1$ can be viewed as a short exact sequence, to get a sense of how many nonequivalent sequences exist we study the $k$-vector space rank of $\operatorname{Ext}^1.$ As before, let $T_1=S/ (x+by+cz)$ and $N=S/ (x+dy+fz).$ If $$\mathbb{F}_N: \qquad \rightarrow S \xrightarrow{\partial_2} S \xrightarrow{\partial_1} S \rightarrow 0$$ is a (deleted) free resolution of $N$, then the complex $\operatorname{Hom}_S(\mathbb{F}_{N}, T_1)$ has the form $$ \qquad 0 \rightarrow T_1 \xrightarrow{\operatorname{Hom}_S(\partial_1, T_1)} T_1\xrightarrow{\operatorname{Hom}_S(\partial_2, T_1)} T_1 \xrightarrow{\operatorname{Hom}_S(\partial_3, T_1)]} T_1 \rightarrow \cdots$$ and $\operatorname{Ext}_S^1(N, T_1)= \operatorname{H}^1(\operatorname{Hom}_S(\mathbb{F}_{N}, T_1))=\ker(\operatorname{Hom}_S(\partial_2, T_1))/\operatorname{im} (\operatorname{Hom}_S(\partial_1, T_1)).$ We can compute the kernels and images of these maps: \begin{eqnarray*}\ker (\operatorname{Hom}_S(\partial_2, T_1)) &=& \left\{ \begin {array}{l l} (\overline{1}) \cong T_1, \quad & \mbox{ if } b=-d \mbox{ and } c=-f\\ (\overline{y}, \overline{z}), \quad & \mbox{ otherwise}\\ \end{array} \right.\\ & & \\ \operatorname{im} (\operatorname{Hom}_S(\partial_1, T_1))&=&((b-d)\overline{y}+ (c-f) \overline{z}). \end{eqnarray*} \noindent Therefore, $$\operatorname{Ext}_S^1(N, T_1)=\left\{ \begin {array}{l l} T_1/((d-b)\overline{y}+ (f-c) \overline{z}), \quad & \mbox{ if } b=-d \mbox{ and } c=-f\\ (\overline{y}, \overline{z})/((d-b)\overline{y}+ (f-c) \overline{z}), \quad & \mbox{ otherwise,}\\ \end{array} \right. $$ and, provided char$(k)\neq 2,$ we have $$\operatorname{rank} (\operatorname{Ext}_S^1(N, T_1))=\left\{ \begin {array}{l l} 3, \quad & \mbox{ if } b=c=d=f=0 \\ 2, \quad & \mbox{ if } b=-d \mbox{ and } c=-f, \mbox{or if } b=d \mbox{ and } c=f\\ 1, \quad & \mbox{ otherwise.}\\ \end{array} \right. $$ Suppose $(u, v)$ are an exact pair of zero divisors. Then for some $f_1,f_2 \in k$, we have $u=x+f_1y+f_2z$ and $v=x-f_1y-f_2z.$ From the above computations, for some $[\alpha] \in \operatorname{Ext}_S^1(S/u,S/v)$ and some $S$-module $M,$ there exists the short exact sequence $$\lambda:0 \rightarrow S/v \rightarrow M \rightarrow S/u \rightarrow 0,$$ where $M$ has a presentation matrix $$\mathbf{M}=\left[\begin{array}{c c} v & -\alpha\\ 0 & u \end {array} \right].$$ In this case, $1_S$ is a possibility for $-\alpha$ since $\overline{1} \in \ker(\operatorname{Hom}_S(\partial_2, S/v)).$ However $$\operatorname{coker} \mathbf{M}= \operatorname{coker} \left[\begin{array}{c c} v & 1_S\\ 0 & u \end {array} \right] \cong S.$$ This implies that $\lambda$ represents part of the complete resolution of $R/v$. Since this is always the case when we have a pair of exact zero divisors, we exclude it to focus on when the non-syzygy cases occurs. Hence, we define $\Gamma_S(N, T_1)$ to represent the rank of the elements in $\operatorname{Ext}_S^1(N, T_1)$ which are not part of a complete resolution of $T_1.$ Therefore, \\ \begin {equation} \label{rank} \Gamma_S(N, T_1) :=\left\{ \begin {array}{l l} 2 \quad & \mbox{ if } b=c=d=f=0 \mbox{ or } b=d \mbox{ and } c=f, \\ 1 \quad & \mbox{ otherwise.}\\ \end{array} \right.\end{equation} \subsection{Bounds on the Ranks of $\operatorname{Ext}_R^1(T_{i}/T_{i-1}, T_{i-1})$} Let $T_n$ be a totally reflexive module over $(R, \mathfrak{m}, k)$ with $\mathfrak{m}^3=0$ that has an $n\times n$ upper triangular minimal presentation matrix. From Theorem 1.4 in \cite{BT}, we know that if there exists one nontrivial totally reflexive modules and if $k$ is of characteristic zero, then there are infinitely many non-isomorphic indecomposable totally reflexive modules of each admissible length, specifically a multiple of the embedding dimension of the ring. Assume that $R$ is also a finite dimension $k$-algebra. We investigate the rank of $\operatorname{Ext}_R^1(T_{i}/T_{i-1}, T_{i-1})$ to get a sense of the complexity (or simplicity) of modules of each possible length. Consider the short exact sequence $$0 \rightarrow T_{i-1} \rightarrow T_{i} \rightarrow T_{i}/T_{-1} \rightarrow 0. \label{T2seq}$$ From this short exact sequence we can obtain a long exact sequence of Ext $$ \cdots \rightarrow \operatorname{Ext}_R^1(T_{i}/T_{i-1}, T_{i-1}) \rightarrow\operatorname{Ext}_R^1(T_{i}/T_{i-1}, T_{i}) \rightarrow \operatorname{Ext}_R^1(T_{i}/T_{i-1}, T_{i}/T_{i-1})\rightarrow \cdots.$$ We then have the inequality \\ \begin{center} \small{ $\operatorname{rank}(\operatorname{Ext}_R^1(T_{i}/T_{i-1}, T_{i})) \leq \operatorname{rank}(\operatorname{Ext}_R^1(T_{i}/T_{i-1},T_{i}/T_{i-1})) + \operatorname{rank}(\operatorname{Ext}_R^1(T_{i}/T_{i-1}, T_{i-1})).$\\} \end{center} Now we will find an upper bound for $ \Gamma(T_{i}/T_{i-1}, T_{i-1})$ over the ring $S=k[X, Y, Z]/(X^2, Y^2, Z^2, YZ).$ By (\ref{rank}), if we consider the case when $i=2,$ then we also have that $1 \leq \Gamma(T_{2}/T_1, T_{1})\leq 2$. Therefore $$\operatorname{rank}\Gamma_S(T_{2}/T_1, T_{2}) \leq 4.$$ By induction we see that $$\Gamma_S(T_{i}/T_{i-1}, T_{i}) \leq 2n \qquad \mbox { for } i=2, 3 \ldots.$$ In fact, we know of cases when this rank is bounded below by one. That is, there exists a nontrivial short exact sequence beginning in $T_i$ and ending in $T_{i}/T_{i-1}.$ To find such a case, per \cite{BT}, we define the $b \times b$ upper triangular matrix \begin{equation} \label{BTmatrix} M_b(s, t, u, v)= \left[\begin{array} {c c c c c c} s & u& 0& 0& 0 & \dots\\ 0& t & v& 0& 0& \dots\\ 0& 0& s & u& 0& \dots \\ 0& 0 & 0& t & v& \\ 0& 0 & 0& 0& s& \ddots \\ \vdots& \vdots& \vdots& \vdots& &\ddots \end{array} \right],\end{equation} and consider the following theorem. \begin{theorem} \label{BTthm}\cite[Theorem 3.1]{BT} Let $(R, \mathfrak{m})$ be a local ring and assume that $ s$ and $ t$ are elements in $\mathfrak{m}\backslash \mathfrak{m}^2 $ that form an exact pair of zero divisors. Assume further that $ u $ and $ v $ are elements in $\mathfrak{m}\backslash \mathfrak{m}^2$ with $ uv=0$ and that one of the following conditions holds: \noindent(a) The elements $s, t,$ and $ u$ are linearly independent modulo $\mathfrak{m}^2.$\\ (b) One has $ s\in (t) +\mathfrak{m}^2 $ and $ u, v \notin (t) + \mathfrak{m}^2.$ \noindent For every $b \in \mathbb{N}$, the $R$-module $M_b(s, t, u, v)$ is indecomposable, totally reflexive, and non-free. Moreover $\operatorname{coker} M_b(s, t, u, v)$ has constant Betti numbers, equal to $b.$ \end{theorem} Over $S$ the all of the exact pairs of zero divisors are of the form $( x+\alpha y+ \beta z, x-\alpha y- \beta z) $ for $ \alpha, \beta \in k$ \ref{expairsTR}. Now consider the choices for $u, v \in \mathfrak{m} \backslash \mathfrak{m}^2$ whose product is zero. From part (a) of Theorem \ref{BTthm}, $ x+\alpha y+ \beta z, x-\alpha y- \beta z $ and $u$ must be linearly independent in $\mathfrak{m}/\mathfrak{m}^2. $ For $ \gamma, \eta, \lambda, \tau \in k $ let $u=\gamma y+\eta z$ and $v= \lambda y + \tau z.$ Therefore, for $\gamma\neq \pm 2 \alpha$ and $\eta \neq \pm 2 \beta$ we have that $M_b(x+\alpha y+ \beta z, x-\alpha y- \beta z, \gamma y+\eta z, \lambda y + \tau z)$ is a presentation matrix of a nontrivial indecomposable totally reflexive $S$-module. Although Theorem \ref{BTthm} is useful in finding some of the indecomposable totally reflexive modules that have an upper triangular presentation matrix, it says nothing about whether the choices of $ s, t , u, $ and $ v$ will lead to non-isomorphic modules. In fact, notice that for the choices above of $u$ and $v$ over $S,$ neither one of them contains an $x$ term. This is because if either of them did, then one can always find an equivalent presentation matrix, and thus an isomorphic module, which does not have an $x$ term on the super diagonal. \section{Over Finite Fields} Let us now consider the ring $\mathbb{Z}_2[X, Y, Z]/(X^2, Y^2, Z^2, YZ).$ Now we can list the four one-generated totally reflexive modules: $$ (x),\quad (x+y),\quad (x+z),\quad (x+y+z),$$ all of which are non-isomorphic. From here we can construct all possible totally reflexive modules that have a $2 \times 2$ upper triangular presentation matrix, say \begin{equation}\left[\begin{array}{c c} u& a\\0&t \end{array} \right]. \label{2by2} \end{equation} However, there may be many modules that have isomorphic presentation matrices. Recall from \cite[Theorem 4.3]{equivmat}, that two modules are isomorphic if and only if they have equivalent presentation matrices. To discover which ones do, we start with the assumption that $a$ does not contain an $x$ term. Since $u$ and $t$ both must have an $x$ term, we have that $\left[\begin{array}{c c} u& a\\0&t \end{array} \right] $ is equivalent to $\left[\begin{array}{c c} u& a-t\\0&t \end{array} \right]. $ Define $$\mathcal{T}=\{x, x+y, x+z, x+y+z\}\qquad \mbox{ and } \qquad \mathcal{ N}=\{y, z, y+z\}, \label {ordering}$$ and hence $u, t \in \mathcal{T}$ and $a \in \mathcal{N}.$ We will also impose an ordering on the elements in $\mathcal{T} $ and $\mathcal{ N}$ as the order listed above from smallest to largest. For $i=1,2$ let $u_i, t_i \in \mathcal{T}$ and $a_i \in \mathcal{N},$ the for two matrices $$M_1=\left[\begin{array}{c c} u_1& a_1\\0&t_1 \end{array} \right] \qquad \mbox{ and } \qquad M_2=\left[\begin{array}{c c}u_2& a_2\\0&t_2 \end{array} \right]$$ define $M_1$ to be \emph{smaller} than $M_2$ if \begin{eqnarray*} &&u_1<u_2 \mbox{ \: or }\\ &&u_1=u_2 \mbox{ \: and \,}t_1<t_2 \mbox{ \: or}\\ &&u_1=u_2,\, t_1=t_2 \mbox{ \: and } a_1<a_2. \end{eqnarray*} Through the use of the CAS Magma, we can find the isomorphism classes of all totally reflexive modules that have a $2 \times 2$ upper triangular presentation matrix. We chose the smallest presentation matrix to represent each class. There are 24 non-isomorphic indecomposable totally reflexive modules with an upper triangular presentation matrix. In the below table, we list representatives for each isomorphism class. For a matrix in the form of (\ref{2by2}), the options for $a$ which represent non-isomorphic indecomposable totally reflexive modules are listed in the center of the table. $$ \begin{array}{| l | | r | r | r | r |} \hline u \; \downarrow \qquad t \ \rightarrow & x & x+y & x+z & x+y+z\\ \hline \hline x & y, z, y+z & z & y&y\\ \hline x+y& z& y, z, y+z & y&y\\ \hline x +z& y & y& y, z, y+z & z\\ \hline x +y+z& y & y& z& y, z, y+z\\ \hline \end{array} $$ Something to note about these modules when the coefficient field is $\mathbb{Z}_2$ is that it is not possible to interchange $u$ and $t,$while keeping the same $a,$ and have them be isomorphic to each other. However, this is not the case if instead we consider modules over $\mathbb{Z}_3[x, y, z]/(x^2, y^2, z^2, yz).$ Over this ring, a module with presentation matrix $\left[\begin{array}{c c} u& a\\0&t \end{array} \right]$ is isomorphic to one with $\left[\begin{array}{c c}t& a\\0&u \end{array} \right]$ as a presentation matrix. Theorem \ref{thm1} is useful in determining how the totally reflexive submodules of a totally reflexive module are contained in one another in this special case. However, there are totally reflexive modules which do not have an upper triangular presentation matrix and thus do not have a saturated TR-filtration. We conclude with an example of such a module. \section{An example with no upper triangular presentation matrix} \begin{example}\label{notUT} Consider the $S$-module $T$ with a presentation matrix of $$\mathbf{T}:=\left[ \begin{array}{cc} x & z \\ y & x \end{array} \right], $$ where $S$ is the same ring defined previously, with char$(k)=0$. This is a totally reflexive module. Take a free resolution of it $$\mathbb{F}_T: \qquad \cdots \rightarrow S^2 \xrightarrow{\left[\begin{array}{c c} x&z\\y&x \end{array}\right]} S^2 \xrightarrow{\left[\begin{array}{l l} x&-z\\-y&x \end{array}\right]} S^2 \xrightarrow{\left[\begin{array}{c c} x&z\\y&x \end{array}\right]} S^2 \rightarrow 0$$ and apply the functor $\operatorname{Hom}_S(\_,S)$. $$ \operatorname{Hom}_S(\mathbb{F}_T, S): \quad 0 \rightarrow S^2 \xrightarrow{\left[\begin{array}{c c} x&z\\y&x \end{array}\right]} S^2 \xrightarrow{\left[\begin{array}{l l} x&-z\\-y&x \end{array}\right]} S^2 \xrightarrow{\left[\begin{array}{c c} x&z\\y&x \end{array}\right]} S^2 \rightarrow \cdots$$ Therefore, $\operatorname{Hom}_S(\mathbb{F}_T, S)\cong \mathbb{F}_T$ and hence the complex $\operatorname{Hom}_S (\mathbb{F}_T, S)$ is also exact. This implies that $T$ is a totally reflexive $S$-module. In particular, this module is not isomorphic to a totally reflexive module that has an upper triangular presentation matrix. This can be done by showing that the matrix $\mathbf{T}$ is not equivalent to an upper triangular matrix. \begin{proposition} For $\mathbf{T} =\left[\begin {array} {c c} x&z\\y&x \end{array} \right], $ a totally reflexive $S$-module, the totally reflexive module $T = \operatorname{coker}(\mathbf{T})$ is not isomorphic to a totally reflexive module that has a upper triangular presentation matrix. \end{proposition} \begin{proof} Suppose that for an upper triangular matrix $\mathbf{U}$ we have $\operatorname{coker} \mathbf{T} \cong \operatorname{coker}\mathbf{ U},$ and hence $\mathbf{T}$ would be equivalent to $\mathbf{U}$. That is, there would exist two invertible matrices $\left[ \begin {array} {c c} a&b\\c&d \end{array} \right]$ and $ \left[ \begin {array} {c c} e&f\\g&h \end{array} \right] $ where $a, b, \ldots, h \in k$ such that \begin{eqnarray} \left[ \begin {array} {c c} a&b\\c&d \end{array} \right] \;\mathbf{T} \; \left[ \begin {array} {c c} e&f\\g&h \end{array} \right] = \mathbf{U}\label{product}. \end{eqnarray} For $\alpha, \beta, \gamma, \zeta, \varphi, \psi, $ and $ \lambda $ in $k,$ let $$ \mathbf{U}:= \left[ \begin {array} {c c} x+\alpha y + \beta z & \zeta x+ \varphi y + \psi z \\ 0 & x+ \gamma y +\lambda z\end{array} \right]$$ \noindent Computing the products in (\ref{product}) gives us \begin{eqnarray*} \left[ \begin {array} {c c} a&b\\c&d \end{array} \right] \left[ \begin {array} {c c} x&z\\y&x \end{array} \right ] \left[ \begin {array} {c c} e&f\\g&h \end{array} \right] &=& \left[ \begin {array} {c c} x+\alpha y + \beta z & \zeta x+ \varphi y + \psi z \\ 0 & x+ \gamma y +\lambda z\end{array} \right] \\ &&\\ &&\\ \small{\left[ \begin{array} {c c}(ae+bg)x +bey+agz & (af+bh)x + bfy +ah z\\ (ce+dg)x +edy+cgz & (cf-dh) x + dfy+chz \end{array}\right]} &=& \left[ \begin {array} {c c} x+\alpha y + \beta z & \zeta x+ \varphi y + \psi z \\ 0 & x+ \gamma y +\lambda z\end{array} \right]. \\ \end{eqnarray*} These yield a system of equations that one can show to be inconsistent. Therefore, $\mathbf{T}$ is not equivalent to $\mathbf{U}$ and thus $T$ is not isomorphic to another totally reflexive module that has a upper triangular presentation matrix. \end{proof} \end{example} \begin{comment} Showing the system of equations is inconsistent. \begin{eqnarray} &&ae+bg=1, \qquad be =\alpha, \qquad ag=\beta, \label{1}\\ &&af+bh=\zeta, \qquad bf=\varphi, \qquad ah = \psi, \label{2}\\ &&ce=-df, \qquad \: \quad ed=0, \qquad cg=0, \label{3}\\ &&cf+dh=1, \qquad df=\gamma, \qquad ch=\lambda. \label{4} \end{eqnarray} From (\ref{3}) assume that $c=0,$ and by (\ref 4) we have that $dh=1$ and $\lambda=0$. Since $d\neq 0, $ we must have that $ h=\frac{1}{d}.$ By (\ref{3} ), $-dg=0$ which implies that $g=0$ and using (\ref{1}), we have $a=\frac{a}{e}$ for $e\neq 0$. By (\ref{3}), $ed=0$ which implies $ d=0$, a contradiction to the matrix $\left[ \begin {array} {c c} a&b\\c&d \end{array} \right]$ being invertible. If instead we let $g=0,$ then this leads to an argument similar to the above one. \end{comment} \subsection*{Acknowledgments} I would like to thank David Jorgensen for his guidance in researching and writing this paper. I am also grateful to Graham Leuschke for his helpful comments and proofreading skills.
\section{ACE Bounds Under the IV Model} We consider the model in which $X$ and $Y$ are binary, taking values in $\{0,1\}$, while $Z$ takes $K$ states $\{1,\ldots,K\}$. We use the notation $X(z_i)$ to indicate $X(z = i)$, similarly $Y(x_j)$ for $Y(x = j)$. We consider four different sets of assumptions: \begin{longlist}[(iii)] \item[(i)] $Z \ind Y({x}_0),Y({x}_1), X({z}_1),\ldots,X({z}_{K})$; \item[(ii)] $Z \ind Y({x}_0),Y({x}_1)$; \item[(iii)] for $i \in\{1,\ldots,K\}$, $j \in\{0,1\}$, $Z \ind X(z_i),\linebreak[4] Y(x_j)$; \item[(iv)] there exists a $U$ such that $U \ind Z$ and for $j \in\{ 0,1\}$, $Y(x_j) \ind X,Z \mid U$. \end{longlist} Condition (i) is joint independence of $Z$ and all potential outcomes for $Y$ and $X$. (ii) does not assume independence (or existence) of counterfactuals for $X$. (iii)~is a subset of the independences in (i), none of which involve potential outcomes from different worlds.\footnote{In other words, they do not involve both $Y(x_0)$ and $Y(x_1)$, nor $X(z_i)$ and $X(z_j$) for $i\neq j$.} The counterfactual independencies (i), (ii), (iii) arise most naturally in the context where the instrument is randomized, as depicted by the DAG in Figure~\ref{figswig}(a). Assumption (iii) may be read (via d-separation) from the Single-World Intervention Graph (SWIG)\footnote{See \citet{richardsonrobins2013} for details.} $\mathcal{G}_1(z,x)$, depicted in Figure~\ref{figswig}(b), which represents the factorization of $P(Z,X(z),Y(x),U)$, implied by the IV model. Lastly (iv) consists of only three independence statements, but does assume the existence of an unobserved variable $U$ that is sufficient to control for confounding between $X$ and $Y$. No assumption is made concerning the existence of counterfactuals $X(z)$; confounding variables ($U^*$) between $Z$ and $X$ are permitted (so long as $U^* \ind U$). The DAG $\mathcal{G}_2$ and corresponding SWIG $\mathcal{G}_2(x)$ are shown in Figure~\ref{figswig}(c), (d). In \citet{richardsonrobins2014}, we prove the following. \begin{thm} Under any of the assumptions \textup{(i)}, \textup{(ii)}, \textup{(iii)}, \textup{(iv)}, the set of possible joint distributions $P(Y(x_0), Y(x_1))$ are characterized by the $8K$ inequalities:\vspace*{-2pt} \begin{eqnarray} \label{eqmarg}&&P\bigl(Y(x_i) = y\bigr) \nonumber\\[-1pt] &&\quad \leq P(Y = y, X = i | Z = z)\\[-1pt] &&\qquad {}+ P(X = 1-i | Z = z),\nonumber \\[4pt] \label{eqjoint}&&P\bigl(Y(x_0) = y, Y(x_1) = \tilde{y}\bigr) \nonumber\\[-1pt] &&\quad \leq P(Y = y, X = 0 | Z = z)\\[-1pt] &&\qquad {} + P(Y = \tilde{y}, X = 1 | Z = z).\nonumber \end{eqnarray} \end{thm} Thus a distribution $P(X,Y | Z)$ is compatible with the stated assumptions if and only if there exists a distribution $P(Y(x_0), Y(x_1))$ satisfying (\ref{eqmarg}) and (\ref{eqjoint}). \begin{thm} Under any of the assumptions \textup{(i)}, \textup{(ii)}, \textup{(iii)}, \textup{(iv)} for all $i,j \in\{0,1\}$, $P(Y(x_i) = j) \leq g(i,j)$, where\vspace*{-2pt} {\fontsize{10.9}{12.9}\selectfont{\begin{eqnarray*} g(i,j) &\equiv&\min \Bigl\{ \min_{z} \bigl[\vphantom{\hat{P}} P(X = i, Y = j | Z = z)\\ &&\hphantom{\min \Bigl\{\min_{z} \bigl[}{} + P(X = 1-i | Z = z) \bigr], \\ &&\hphantom{\min \Bigl\{} \min_{z, \tilde{z}: z \neq\tilde{z}} \bigl[ P(X = i, Y = j | Z = z) \\ &&\hphantom{\min \Bigl\{\min_{z, \tilde{z}: z \neq\tilde{z}} \bigl[}{}+ P(X = 1-i, Y = 0 | Z = z) \nonumber \\ &&\hphantom{\min \Bigl\{\min_{z, \tilde{z}: z \neq\tilde{z}} \bigl[}{} + P(X = i, Y = j | Z = \tilde{z}) \\ &&\hphantom{\min \Bigl\{\min_{z, \tilde{z}: z \neq\tilde{z}} \bigl[}{}+ P(X = 1-i, Y = 1 | Z = \tilde {z}) \bigr] \Bigr\}. \nonumber \end{eqnarray*}}} Furthermore, $P(Y(x_0))$ and $P(Y(x_1))$ are variation independent. Consequently, \begin{eqnarray*} 1-g(1,0)-g(0,1) &\leq& \operatorname{ACE}(X \rightarrow Y)\\ & \leq& g(0,0)+g(1,1)-1. \end{eqnarray*} These bounds are sharp. \end{thm} \begin{figure} \includegraphics{485f01.eps} \caption{\textup{(a)} IV model with no confounding between $Z$ and $X$; \textup{(b)} SWIG representing $P(Z, X(z),Y(x),U)$; \textup{(c)} IV model with confounding between $Z$ and $X$; \textup{(d)} SWIG representing $P(Z, X,Y(x),U,U^*)$.}\label{figswig} \end{figure} Note that to evaluate $g(i,j)$ requires finding a minimum over $K^2$ expressions. In the case where $K=2$, these bounds reduce to those given by \citet {BalkPearboun1997}, who assume (i).\footnote{\citet{dawid2003} working in a non-counterfactual framework also established the bounds for $K=2$ under the DAG in Figure~\ref{figswig}(a); however, his proof also applies to Figure~\ref{figswig}(c). \citet {robinsgreenland1996} observed that the Balke--Pearl bounds were also sharp under (ii).} \citet{robins1989} and \citet{manski1990} derived what are called the ``natural bounds'' on the ACE under the weaker assumption that $Z \ind Y({x}_0)$ and $Z \ind Y({x}_1)$. As noted by Imbens, without further assumptions these bounds are not sharp. However, the natural bounds are sharp under (i) or (iii), if, in addition, we assume there are no Defiers (an assumption that has testable implications). \citet{chengsmall2006} considered bounds on the ACE when $K=3$ under additional assumptions. \section{Market Equilibrium and BiCausal Models} Imbens' clear description of the market equilibrium model is particularly informative. We also strongly endorse the author's contention that the RHS of systems of structural equations should be interpreted as describing potential outcomes for the LHS.\footnote{ \citet{pearl2000}, \citet{laucausal}, \citet{lauritzen02} argue that these are not really ``equations'' but are better viewed as ``assignments'' in computer languages, for example, $ y \leftarrow x +1$; see also \citet{strotzwoldrecursive1960}, page 420.} However, we note that this position has important implications both for interpretation and inference. Furthermore, it does not seem to be universally accepted within Economics. \citet{leroy2006} states that ``economic models use the equality symbol with its usual mathematical meaning, not with the meaning of the assignment operator''; an approach that is clearly incompatible with an interpretation in terms of potential outcomes. For example, it becomes permissible to renormalize structural equations to change which variable is on the LHS. It has also been argued that statistical analyses of such models should be invariant to the normalization; see \citet{hillier1990}, \citet{basmanncausal1963}.\hskip.2pt\footnote{For example, \citet{greene2003}, page 401, states (in the context of the IV model): ``one significant virtue of [the Limited Information Maximum Likelihood Estimator] is its invariance to normalization of the equations.''} Contrary to Imbens' remark,\footnote{Footnote 8, page 331.} this alternative view does not appear to be motivated by considerations of measurement error. \citet{leroy2006} makes clear that he does not believe that structural equations describe potential outcomes for endogenous variables and does not discuss issues relating to measurement.\footnote {For example, \citet{leroy2006}, page 23, states that ``The assumption that it makes sense to delete one or more of the structural equations and replace the value of the internal variable so determined by a constant without altering the other equations [\ldots] is virtually never satisfied in economic models since each external variable typically affects equilibrium values of more than one internal variable.'' He goes on to assert ``In fact, it is difficult to think of nontrivial models in any area of research in which the [\ldots] assumption is satisfied.''} Rather, this appears to be a fundamental difference in interpretation. The market equilibrium model specifies potential outcomes for $Q^d_t(p)$, $Q^s_t(p)$: \begin{eqnarray} \label{eqqd}Q^d_t(p)&=& \alpha^d + \beta^d p + \varepsilon_t^d, \\ \label{eqqs}Q^s_t(p)&=& \alpha^s + \beta^s p + \varepsilon_t^s, \end{eqnarray} and imposes the equilibrium condition:\footnote{To simplify notation, throughout we work directly in terms of $\log$ price and $\log$ quantity.} \begin{eqnarray} Q^d_t(p) = Q^s_t(p).\label{eqequ} \end{eqnarray} \citet{strotzwoldrecursive1960} described such systems as \textit {bicausal}. It should be observed that the model does not specify potential outcomes for price ($P_t(q_s,q_d)$), nor does it view price as externally determined (i.e., exogenous). Instead price is determined implicitly as a consequence of the equilibrium condition. In this regard, the model might be regarded as incomplete: Indeed \citet {haavelmowhat1958} is quite critical of this model for failing to offer any \emph{explanation} as to how the equilibrium price is determined. The model also falls outside the scope of non-parametric structural equation models (NPSEM) (see, e.g., \cite{pearl2000}), which require one equation for each endogenous variable;\footnote Indeed \citet{leroy2006} argues against the interpretation of structural equations in terms of potential outcomes on the grounds that this interpretation, as advanced by Pearl, requires a one-to-one mapping between equations and endogenous variables that he argues, does not make sense for the market equilibrium model.} likewise the model defies standard graphical representation, though see Figure~\ref{figone}(a). \begin{figure} \includegraphics{485f02.eps} \caption{\textup{(a)} Attempt to depict the bicausal model; \textup{(b)} a schematic showing the deterministic system (\protect\ref{eqcon})--(\protect\ref{eqmer}); the edge \protect\tikz\protect\path(0ex,0ex) edge[->] node[above=0pt, black] {$\scriptscriptstyle I$} (3ex,0ex); denotes that $P$ is the integral of $\Delta P$; see Iwasaki and Simon (\citeyear {iwasaki1994}).}\label{figone} \end{figure} A related question concerns whether there exist dynamic acyclic (i.e., recursive) systems of structural equations that lead to the equilibrium distribution corresponding either to a cyclic system of structural equations or a bicausal system.\footnote Analysis of this question was stimulated by a heated debate that arose between Wold, who advocated a recursive, regression-based approach to demand analysis, and Haavelmo and the Cowles Commission who advocated simultaneous equations. See \citet{haavelmostatistical1943}, \citet{woldbentzelstatistical1946}, \citet{woldjureendemand1953}, \citet{bentzelhansenrecursiveness1954}, \citet{strotzwoldrecursive1960}, \citet{basmanncausal1963}; historical overviews are given by \citet{morganstamping1991}, \citet{epsteinhistory1987}.} \citet{fishcorr} provides just such a ``correspondence principle'' under which the distribution implied by a cyclic linear SEM is obtained as a time average of a deterministic set of first order difference equations reaching a static equilibrium subject to stochastic boundary conditions. The correspondence assumes that the equilibration time is very fast relative to the interval between observations so the time averaged variables are in deterministic equilibrium. Fisher also derived conditions on the coefficient matrices of a cyclic SEM that are required in order for the system to reach equilibrium; in fact he further required that each subset of structural equations also have this property. However, Fisher's correspondence presumes a normalization under which each variable is associated with a single equation (as in an NPSEM), and hence would not apply to a bicausal system. \citet{richphd}, Chapter~2, described a system of finite difference equations that gives rise to the bicausal system~(\ref{eqqd})--(\ref{eqequ}): \begin{eqnarray} \label{eqcon}\mbox{Consumers:}&&\hspace*{4pt} Q^d_{t+(k+1)\delta}(p_{t+k\delta})\nonumber \\[-8pt]\\[-8pt] &&\hspace*{4pt}\quad = \alpha^d+ \beta^d p_{t+k\delta} + \varepsilon_{t}^d,\nonumber \\ \label{eqsup}\mbox{Suppliers:}&&\hspace*{4pt} Q^s_{t+(k+1)\delta}(p_{t+k\delta}) \nonumber\\[-8pt]\\[-8pt] &&\hspace*{4pt}\quad = \alpha^s+ \beta^s p_{t+k\delta} + \varepsilon_{t}^s,\nonumber \\ \label{eqmer}\mbox{Merchants:}&&\hspace*{4pt} P_{t+(k+1)\delta}\bigl(q^d_{t+k\delta}, q^s_{t+k\delta},p_{t+k\delta}\bigr) \nonumber\\[-8pt]\\[-8pt] &&\hspace*{4pt}\quad = p_{t+k\delta} + \lambda \bigl(q^d_{t+k\delta} - q^s_{t+k\delta} \bigr),\nonumber \end{eqnarray} for $k=\{0,\ldots, \delta^{-1}-1\}$. Note that the disturbances $(\varepsilon_{t}^d, \varepsilon_{t}^s)$ represent boundary conditions and hence remain fixed during the interval $[t,t+1)$. As in Fisher's correspondence, the observed variables correspond to limiting time-averages over a unit interval: \begin{eqnarray*} \overline{Q}^d_t &=& \lim_{\delta\rightarrow0} \delta \sum_{k=0}^{\delta^{-1}-1} {Q}^d_{t+k\delta},\quad \overline{Q}^s_t = \lim_{\delta\rightarrow0} \delta \sum_{k=0}^{\delta^{-1}-1} {Q}^s_{t+k\delta},\\ \overline{P}_t &=& \lim_{\delta\rightarrow0} \delta\sum _{k=0}^{\delta^{-1}-1} {P}_{t+k\delta}. \end{eqnarray*} Under suitable conditions on the coefficients, $(\overline{Q}^d_t, \overline{Q}^s_t,\allowbreak \overline{P}_t)$ obey equations (\ref{eqqd})--(\ref{eqequ}). Note that Merchants' equation (\ref{eqmer}) which includes $P$, leads to the equilibrium condition (\ref{eqequ}) that does not.\footnote{In causal terms, this model is similar to one presented in \citet{wold1959}. Wold viewed his model as a formalization of Cournot's theories.} It might be objected to the proposed model that there is no disturbance term in equation (\ref{eqmer}). The explanation for this is that the disturbance terms in the nonrecursive model correspond to constant factors in the deterministic evolution. The equation for price gives the change in price during a small interval (length $\delta$) to the discrepancy between supply and demand. Adding a disturbance term would say that throughout the observation period (length $1$) the Merchants' reaction to change in price was off by a constant factor, so that even if quantities supplied and demanded were identical, the Merchants would change the price. Thus, if we add an error $\varepsilon^p_t$ the model will not, in general, arrive at equilibrium within the unit interval.\footnote{Having said this, the equations (\ref{eqcon}) and (\ref{eqsup}) still imply that producers and consumers make systematic errors in computing prices over a time-scale of length $\delta$.} \citet{iwasaki1994} represent equilibrating mechanisms via ``causal influence diagrams'' in which the derivatives of variables are included. Under this scheme, model (\ref{eqcon})--(\ref{eqmer}) is represented by the graph in Figure~\ref{figone}(b). This example serves to show that time averages of (deterministic) equilibrating systems need not have a structural equation for each variable. See also \citep{2001.dash.esqaru} for related work. \section*{Acknowledgments} This work was supported by the US National Institutes of Health Grant R01 AI032475; Richardson was also supported by the US National Science Foundation Grant CNS-0855230.
\section{Introduction} Consider the second-order diffusion equation \begin{equation} \label{eq:pde} \begin{aligned} -\nabla\cdot (\mathbb{A}\nabla u) &= f \qquad \textrm{in }\Omega,\\ u &= 0 \qquad \textrm{on } \partial\Omega. \end{aligned} \end{equation} where $\Omega$ is a polygonal or polyhedral domain in $\mathbb{R}^d\; (d=2,3)$. Assume that $\mathbb{A}$ is a symmetric, uniformly positive definite, and uniformly bounded-above diffusion matrix. Namely, there exist positive constants $\alpha$ and $\beta$ such that \begin{equation}\label{matrix} \alpha\xi^T\xi\leq \xi^T \mathbb A(x)\xi\le \, \beta\xi^T\xi \quad \text{for all } \xi\in \mathbb{R}^d\;\textrm{and}\; x\in\Omega. \end{equation} The goal of this paper is to construct and analyze an auxiliary space multigrid preconditioner for the weak Galerkin finite element discretization of Problem \eqref{eq:pde}. The weak Galerkin method was recently introduced in \cite{WangYe_PrepSINUM_2011} for second order elliptic equations. It is an extension of the standard Galerkin finite element method where classical derivatives were substituted by weakly defined derivatives on functions with discontinuity. Optimal order of {\em a priori} error estimates has been observed and established for various weak Galerkin discretization schemes for second order elliptic equations \cite{mwy-wg-stabilization, WangYe_PrepSINUM_2011, wy-mixed}. An {\em a posteriori} error estimator was given in~\cite{Chen2013}. Numerical implementations of weak Galerkin were discussed in \cite{mwy-wg-stabilization, MuWangWangYe} for some model problems. The weak Galerkin method has already demonstrated many nice properties in various cases~\cite{WG-biharmonic, mwy-wg-stabilization, WangYe_PrepSINUM_2011, wy-mixed}. Thus we are motivated to study fast solvers and preconditioning techniques for the weak Galerkin method. The main results of this paper are: \begin{itemize} \item We develop a fast auxiliary space preconditioner for weak Galerkin methods using Raviart-Thomas element and Brezzi-Douglas-Marini element on triangular grids. \item We consider both the original system and the reduced system. The original weak Galerkin discretization of \eqref{eq:pde} involves degrees of freedom both on the interior of each mesh element and on mesh edges/faces. The reduced system, which only involves degrees of freedom on edges/faces, is to our knowledge first rigorously constructed and analyzed here. \end{itemize} Recently, Li and Xie announced an auxiliary space mulrigrid preconditioning method for the weak Galerkin finite element method, and the result was posted on ArXive~\cite{LiXie}. This result became to be known to us after the bulk portion of the present paper was developed. Following a through comparison, we conclude that our results are more general, and the two approaches are different in analysis. In addition, our result offers a new feature by covering the weak Galerkin method in the reduced system. We shall briefly introduce the auxiliary space preconditioner constructed in~\cite{xu}. A classical geometric multigrid method constructs discrete spaces on different mesh levels using the same type of discretization. For example, in the classical multigrid method for $H^1$ conforming piecewise linear ($P_1$) finite element approximation, one uses a set of nested meshes with characteristic mesh sizes $h$, $2h$, $4h$, $\ldots$, from the finest mesh to the coarsest. An illustration of V-cycle multigrid is given in Figure \ref{fig:multigrid}. The auxiliary space multigrid method can be essentially understood as a two-level method involving a ``fine'' level and a ``coarse'' level, while the ``fine'' space and ``coarse'' space are not necessarily using the same type of discretization or the same type of meshes. This gives great freedom in choosing the ``coarse'' space, which is also called an auxiliary space. Here we use the weak Galerkin discretization for the ``fine'' level, and the $H^1$ conforming piecewise linear finite element discretization for the ``coarse'' level. Both the ``fine'' level and the ``coarse'' level are discretized on the same mesh, as shown in Figure \ref{fig:multigrid}. In the figure, we conveniently use black rectangles and black dots to denote different type of discretization spaces on different levels. Because the fast solvers for the $H^1$ conforming piecewise linear finite element discretization have been thoroughly studied, one can use any existing solvers/preconditioners as a ``coarse'' solver. For example, one may use a classical multigrid method as a ``coarse'' solver and consequently achieves a true ``multi''-grid effect (see Figure \ref{fig:multigrid}). \begin{figure} \begin{center} \caption{Illustration of auxiliary space multigrid. We use black rectangles and black dots to denote different type of discretization spaces. The dashed circles shows how to derive an auxiliary space ``multi''-grid method by using a classical multigrid as a coarse solver in the two-grid auxiliary space multigrid framework.} \label{fig:multigrid} \includegraphics[width=10cm]{multigrid} \end{center} \end{figure} The rest of the paper is organized as following. In Section \ref{sec:weakGalerkin}, we give a brief introduction of the weak Galerkin method, and in Section \ref{sec:multigrid}, we construct the auxiliary space multigrid preconditioner for the weak Galerkin discretization, and prove that the condition number of the preconditioned system does not depend on the mesh size. After that, we consider a reduced system of the weak Galerkin discretization in Section \ref{sec:reduced} and construct an auxiliary space multigrid solver/preconditioner the reduced system, again with an optimal condition number estimate. Finally in Section \ref{sec:numerical}, we present supporting numerical results. \section{A Weak Galerkin Finite Element Scheme} \label{sec:weakGalerkin} In this section, we give a brief introduction to the weak Galerkin method. Related notation, definitions, and several important inequalities will be stated. Let $D\subseteq\Omega$ be a polygon or polyhedron, we use the standard definition of Sobolev spaces $H^s(D)$ and $H_0^s(D)$ with $s\ge 0$ (e.g., see \cite{adams, ciarlet} for details). The associated inner product, norm, and semi-norms in $H^s(D)$ are denoted by $(\cdot,\cdot)_{s,D}$, $\|\cdot\|_{s,D}$, and $|\cdot|_{r,D}, 0\le r \le s$, respectively. When $s=0$, $H^0(D)$ coincides with the space of square integrable functions $L^2(D)$. In this case, the subscript $s$ is suppressed from the notation of norm, semi-norm, and inner products. Furthermore, the subscript $D$ is also suppressed when $D=\Omega$. For $s<0$, the space $H^s(D)$ is defined to be the dual of $H_0^{|s|}(D)$. The above definition/notation can easily be extended to vector-valued and matrix-valued functions. The norm, semi-norms, and inner-product for such functions shall follow the same naming convention. In addition, all these definitions can be transferred from a polygonal/polyhedral domain $D$ to an edge/face $e$, a domain with lower dimension. Similar notation system will be employed. For example, $\|\cdot\|_{s,e}$ and $\|\cdot\|_e$ would denote the norm in $H^s(e)$ and $L^2(e)$ etc. We also define the $H({\rm div})$ space as follows $$ H({\rm div},\Omega) = \{{\bf q}:\ {\bf q} \in [L^2(\Omega)]^d,\: \nabla\cdot{\bf q}\in L^2(\Omega)\}. $$ Using the notation defined above, the variational form of Equation \eqref{eq:pde} can be written as: Given $f\in L^2(\Omega)$, find $u\in H_0^1(\Omega)$ such that \begin{equation} \label{eq:weakformulation} (\mathbb A \nabla u, \, \nabla v) = (f, v)\qquad \textrm{for all } v\in H_0^1(\Omega). \end{equation} It is well known that equation \eqref{eq:weakformulation} admits a unique solution. In addition, we assume that the solution to \eqref{eq:weakformulation} has $H^{1+s}$ regularity~\cite{Brenner, Grisvard85}, where $0<s\le 1$. In other words, the solution $u$ is in $H^{1+s}(\Omega)$ and there exists a constant $C$ independent of $u$ such that \begin{equation} \|u\|_{1+s} \le C\|f\|_0. \end{equation} Next, we present the weak Galerkin method for solving \eqref{eq:weakformulation}. Let ${\cal T}_h$ be a shape-regular, quasi-uniform triangular/tetrahedral mesh on the domain $\Omega$, with characteristic mesh size $h$. For each triangle/tetrahedron $K\in {\cal T}_h$, denote by $K_0$ and $\partial K$ the interior and the boundary of $K$, respectively. Geometrically, $K_0$ is identical to $K$. Therefore, later in the paper, we often identify these two if it causes no ambiguity. The boundary $\partial K$ consists of three edges in two-dimension, or four triangles in three-dimension. Denote by ${\mathcal E}_h$ the collection of all edges/faces in ${\cal T}_h$. For simplicity, throughout the paper, we use ``$\lesssim$'' to denote ``less than or equal to up to a general constant independent of the mesh size or functions appearing in the inequality''. Let $j$ be a non-negative integer. On each $K\in {\cal T}_h$, denote by $P_j(K_0)$ the set of polynomials with degree less than or equal to $j$. Likewise, on each $e\in {\mathcal E}_h$, $P_j(e)$ is the set of polynomials of degree no more than $j$. Following \cite{WangYe_PrepSINUM_2011}, we define a weak discrete space on mesh ${\mathcal T}_h$ by $$ \begin{aligned} V_{h} = \{v:\: v|_{K_0}\in P_j(K_0)\textrm{ for } K\in {\mathcal T}_h; \ v|_e\in P_l(e)\textrm{ for } e\in {\mathcal E}_h, &\\ \textrm{and } v|_e=0 \textrm{ for } e\in {\mathcal E}_h\cap\partial\Omega\}, \text{ where }l = {j} \text{ or } j+1. & \end{aligned} $$ Observe that the definition of $V_h$ does not require any continuity of $v\in V_h$ across interior edges/faces. A function in $V_h$ is characterized by its value on the interior of each mesh element plus its value on edges/faces. Therefore, it is convenient to represent functions in $V_h$ with two components, $v=\{v_0, v_b\}$, where $v_0$ denotes the value of $v$ on all $K_0$ and $v_b$ denotes the value of $v$ on ${\mathcal E}_h$. The polynomial space $P_l(e)$ consists of two choices: $l= j$ or $j+1$ and the corresponding weak function space will sometimes be abbreviated as $W_{j,j}$ or $W_{j,j+1}$, respectively. The weak Galerkin method seeks an approximation $u_h\in V_h$ to the solution of problem \eqref{eq:weakformulation}. To this end, we first introduce a discrete gradient operator, which is defined element-wisely on each $K\in {\mathcal T}_h$. For the choices of $V_h$ given above, i.e., using $W_{j,j}$ or $W_{j,j+1}$, suitable definitions of the weak gradient involve the Raviart-Thomas (RT) element and the Brezzi-Douglas-Marini (BDM) element, respectively. Let $K$ be either a triangle or a tetrahedron and denote by $\widehat P_k(K)$ the set of homogeneous polynomials of order $k$ in the variable ${\bf x}=(x_1,\ldots, x_d)^T$. Define the BDM element by $G_j(K)=\left[P_{j+1}(K)\right]^d$ and the RT element by $G_{j}(K) = \left[P_j(K)\right]^d + \widehat P_j(K) {\bf x}$ for $j\geq 0$. Then, define a discrete space $$ \Sigma_h = \{{\bf q}\in (L^2(\Omega))^d:\: {\bf q}|_K \in G_j(K)\textrm{ for } K\in {\mathcal T}_h\}. $$ Here in the definition of $V_h$ and $\Sigma_h$, the RT element is paired with $W_{j,j}$ while the BDM element is paired with $W_{j,j+1}$. Note that $\Sigma_h$ is not necessarily a subspace of $H({\rm div},\Omega)$, since it does not require any continuity in the normal direction across mesh edges/faces. \begin{definition}[Discrete Weak Gradient] The discrete weak gradient of $v_h$ denoted by $\nabla_{w}v_h$ is defined as the unique polynomial $(\nabla_{w}v_h)|_K \in G_j(K)$ satisfying the following equation \begin{equation}\label{dwg} (\nabla_{w}v_h, {\bf q})_K = -(v_0,\nabla\cdot {\bf q})_K+ \langle v_b, {\bf q}\cdot{\bf n}\rangle_{\partial K}\quad \text{for all } {\bf q}\in G_j(K), \end{equation} where ${\bf n}$ is the unit outward normal on $\partial K$. \end{definition} Clearly, such a discrete weak gradient is always well-defined. Furthermore, if $v\in H^1(K)$, i.e. $v_b = v_0|_{\partial K}$, and $\nabla v\in G_j(K)$, then one has $\nabla _{w}v = \nabla v.$ In this paper we only consider the $W_{j,j}-RT$ and $W_{j,j+1}-BDM$ pairs on simplicial elements in the discretization. But there are many other different choices of discrete spaces in the weak Galerkin method, defined on either simplicial meshes or other types of meshes including general polytopal meshes~\cite{mwy-wg-stabilization}. Extension of the multigrid preconditioner to other weak Galerkin discretizations will be considered in the future work. We define an $L^2$ projection from $H_0^1(\Omega)$ onto $V_h$ by setting $Q_h v \equiv \{Q_0 v,\, Q_b v\}$, where $Q_0 v|_{K_0}$ is the local $L^2$ projection of $v$ to $P_j(K_0)$, for $K\in{\mathcal T}_h$, and $Q_b v|_e$ is the local $L^2$ projection to $P_l(e)$, for $e\in {\mathcal E}_h$. We also introduce $\mathbb Q_h$ the $L^2$ projection onto $\Sigma_h$. It is not hard to see the following operator identity~\cite{WangYe_PrepSINUM_2011}: \begin{equation}\label{eq:QDDQ} \mathbb Q_h \nabla u = \nabla _wQ_h u, \quad \text{ for all } u\in H_0^1(\Omega) \end{equation} For the $W_{j,j}-RT$ and $W_{j,j+1}-BDM$ pairs, it follows directly from \eqref{eq:QDDQ} that the discrete weak gradient is a good approximation to the classical gradient, as summarized in the following lemma~\cite{WangYe_PrepSINUM_2011}: \medskip \begin{lemma} \label{lem:assumptions} For any $v_h=\{v_0,\,v_b\}\in V_h$ and $K\in{\mathcal T}_h$, $\nabla_w v_h|_K = 0$ if and only if $v_0=v_b = constant$ on $K$. Furthermore, for any $v\in H^{m+1}(\Omega)$, where $0\le m\le j+1$, we have $$\|\nabla_w (Q_h v)-\nabla v\|\lesssim h^{m}\|v\|_{m+1}.$$ In particular, for $v\in H^1(\Omega)$, the $L^2$-projection $Q_h$ is energy stable, i.e, \begin{equation}\label{eq:H1stable} \|\nabla _w (Q_h v)\|\lesssim \|\nabla v\| \quad \text{ for } v\in H^1(\Omega). \end{equation} \end{lemma} Now we are able to present the weak Galerkin finite element formulation for \eqref{eq:weakformulation}: Find $u_h = \{u_0,\, u_b\}\in V_h$ such that \begin{equation} \label{eq:wg} a_h(u_h, v_h) = (f, v_0) \qquad \textrm{for all }v_h = \{v_0,\, v_b\}\in V_h, \end{equation} where the bilinear form $a_h(\cdot,\cdot)$ on $V_h \times V_h$ is defined by \begin{equation}\label{eq:bilinearform} a_h(u_h, v_h) := (\mathbb{A} \nabla_w u_h,\, \nabla_w v_h). \end{equation} The well-posedness and error estimates of the weak Galerkin formulation \eqref{eq:wg} have been discussed in~\cite{WangYe_PrepSINUM_2011, WG-biharmonic}. To state these results, we first define a few discrete inner-products and norms. For any $v_h=\{v_0, v_b\}$ and $\phi_h=\{\phi_0, \phi_b\}$ in $V_h$, define a discrete $L^2$ inner-product by $$ {(\hspace{-.03in}(} v_h, \phi_h{)\hspace{-.03in})} \triangleq \sum_{K\in {\mathcal T}_h}\left [(v_0, \phi_0)_K + h ( v_0-v_b, \phi_0-\phi_b)_{\partial K}\right ]. $$ It is not hard to see that ${(\hspace{-.03in}(} v_h, v_h{)\hspace{-.03in})} = 0$ implies $v_h\equiv 0$. Hence, the inner-product is well-defined. Notice that the inner-product ${(\hspace{-.03in}(}\cdot,\cdot{)\hspace{-.03in})}$ is also well-defined for any $v\in H^1(\Omega)$, for which $v_b|_e=v|_e$ is the trace of $v$ on the edge $e$. In this case, the inner-product ${(\hspace{-.03in}(}\cdot,\cdot{)\hspace{-.03in})}$ is identical to the standard $L^2$ inner-product. Define on each $K\in {\mathcal T}_h$ $$ \begin{aligned} \| v_h \|_{0,h,K}^2 &= \|v_0\|_{0,K}^2 + h \|v_0-v_b\|_{\partial K}^2, \\ \| v_h \|_{1,h,K}^2 &= \|v_0\|_{1,K}^2 + h^{-1} \|v_0-v_b\|_{\partial K}^2, \\ | v_h |_{1,h,K}^2 &= |v_0|_{1,K}^2 + h^{-1} \|v_0-v_b\|_{\partial K}^2. \end{aligned} $$ Using the above quantities, we define the following discrete norms and semi-norms on the discrete space $V_h$ $$ \begin{aligned} \| v_h \|_{0,h} &:= \left(\sum_{K\in {\mathcal T}_h}\| v_h \|_{0,h,K}^2 \right)^{1/2},\\ \| v_h \|_{1,h} &:= \left(\sum_{K\in {\mathcal T}_h}\| v_h \|_{1,h,K}^2 \right)^{1/2}, \\ | v_h |_{1,h} &:= \left(\sum_{K\in {\mathcal T}_h}| v_h |_{1,h,K}^2 \right)^{1/2}. \end{aligned} $$ It is clear that $\| v_h \|_{0,h}^2 = {(\hspace{-.03in}(} v_h, v_h{)\hspace{-.03in})}$. Moreover, we point out that the above norms and semi-norms are also well-defined for functions in $H^1(\Omega)$. In this case they are identical to the usual $L^2$-norm, $H^1$-norm, and $H^1$-seminorm, respectively. With the aid of the above defined norms, we state an additional estimate of the $L^2$ projection $Q_h$, which was proved in \cite{WG-biharmonic}. \begin{lemma} \label{lem:Qh} For any $v\in H^{m}(\Omega)$ with $\frac{1}{2} < m \le j+1$, we have \begin{equation} \label{eq:a5} \|v-Q_h v\|_{0,h} \lesssim h^m \|v\|_{m}. \end{equation} \end{lemma} The following three Lemmas have also been proved in~\cite{WG-biharmonic}. First, we have the equivalence between $\|\nabla_w(\cdot)\|$ and the $|\cdot|_{1,h}$ semi-norm: \medskip \begin{lemma} \label{lem:discretenorm-equivalence} For any $v_h=\{v_0, v_b\}\in V_h$, we have \begin{equation} \label{eq:discreteh1semi} |v_h|_{1,h} \lesssim \|\nabla_w v_h\| \lesssim | v_h|_{1,h}. \end{equation} \end{lemma} Moreover, the discrete semi-norms satisfy the usual inverse inequality, as stated in the following Lemma. \medskip \begin{lemma} \label{lem:discreteinverse} For any $v_h=\{v_0, v_b\}\in V_h$, we have \begin{equation} \label{eq:discreteinverse} |v_h|_{1,h} \lesssim h^{-1} \|v_h\|_{0,h}. \end{equation} Consequently, by combining \eqref{eq:discreteh1semi} and \eqref{eq:discreteinverse}, we have \begin{equation} \|\nabla_w v_h\| \lesssim h^{-1} \| v_h\|_{0,h}. \end{equation} \end{lemma} Next, the discrete semi-norm $\|\nabla_w(\cdot)\|$, which is equivalent to $|\cdot|_{1,h}$ as shown in Lemma \ref{lem:discretenorm-equivalence}, satisfies a Poincar\'{e}-type inequality. \medskip \begin{lemma} \label{lem:discretepoincare} The Poincar\'{e}-type inequality holds true for functions in $V_{h}$. In other words, we have the following estimate: \begin{equation} \| v_h \|_{0,h} \lesssim \|\nabla_w v_h\|\qquad \textrm{for all } \ v_h\in V_{h}.\label{eq:discretepoincare1} \end{equation} \end{lemma} Following the above lemmas and \eqref{matrix}, it is clear that equation \eqref{eq:wg} admits a unique solution. This, together with error estimates for the weak Galerkin method, has been proved in~\cite{WangYe_PrepSINUM_2011}. \medskip \begin{theorem} \label{thm:apriori} Assume Problem \eqref{eq:weakformulation} has $H^{1+s}$ regularity, where $0< s\le 1$. The weak Galerkin problem \eqref{eq:wg} admits a unique solution. Let $u\in H_0^1(\Omega)\cap H^{m+1}(\Omega)$, $0\le m\le j+1$, be the solution to \eqref{eq:weakformulation} and $u_h = \{u_{h,0}, u_{h,b}\}$ be the solution to \eqref{eq:wg}, then we have \begin{align} \|\nabla_w (Q_h u - u_h)\| &\lesssim h^m \|u\|_{m+1} , \\ \|Q_0 u - u_{h,0}\| &\lesssim h^{m+s} \|u\|_{m+1} + h^{1+s} \|f-Q_0f\|.\label{eq:L2} \end{align} \end{theorem} \begin{remark} \label{rem:apriori} Theorem \ref{thm:apriori} is only stated for homogeneous Dirichlet boundary value problems. Similar results hold for problems with non-homogeneous Dirichlet boundary or Neumann boundary conditions \cite{WG-biharmonic, WangYe_PrepSINUM_2011}. \end{remark} At the end of this section, we state a scaled trace theorem. Let $K$ be an element with $e$ as an edge. It is well known that for any function $g\in H^1(K)$ one has \begin{equation}\label{trace} \|g\|_{e}^2 \lesssim h^{-1} \|g\|_K^2 + h \|\nabla g\|_{K}^2. \end{equation} \section{An auxiliary space multigrid preconditioner} \label{sec:multigrid} In this section, we construct an auxiliary space multigrid method for the weak Galerkin formulation \eqref{eq:wg}. The auxiliary space multigrid method was introduced by J. Xu in~\cite{xu}. Its main idea is to use an auxiliary space as a ``coarse'' space in the multigrid algorithm, where the discrete problem in the auxiliary space can be easily solved by an existing solver. In our construction, we will use the $H^1$ conforming piecewise linear finite element space as an auxiliary space. The main technical difficulty is to build the connection between the weak Galerkin discrete space $V_h$ and the $H^1$ conforming piecewise linear finite element space. Define the auxiliary space $V^c_h \subset H_0^1(\Omega)$ to be $H^1$ the conforming piecewise linear finite element space on mesh $\mathcal{T}_h$. The spaces $V_h$ and $V^c_h$ are equipped with inner-products ${(\hspace{-.03in}(}\cdot,\cdot{)\hspace{-.03in})}$ and $(\cdot,\cdot)$, and induced norms $\|\cdot\|_{0,h}$ and $\|\cdot\|$, respectively. Define linear operators $A:\, V_h\rightarrow V_h$ and $\mathcal{A}:\: V^c_h\rightarrow V^c_h$ by \begin{equation} \label{eq:Adefinition} \begin{aligned} {(\hspace{-.03in}(} A u,\, v {)\hspace{-.03in})} &= (\mathbb{A} \nabla_w u, \, \nabla_w v) \qquad &&\textrm{for all }v\in V_h, \\ (\mathcal{A} u,\, v) &= (\mathbb{A} \nabla u,\, \nabla v) \qquad &&\textrm{for all }v\in V^c_h. \end{aligned} \end{equation} By the Poincar\'{e} inequality and Lemma \ref{lem:discretepoincare}, it is clear that operators $A$ and $\mathcal{A}$ are symmetric and positive definite with respect to ${(\hspace{-.03in}(}\cdot,\cdot{)\hspace{-.03in})}$ and $(\cdot,\cdot)$, respectively. Hence we can define the $A$-norm and $\mathcal{A}$-norm on $V_h$ and $V^c_h$, respectively, by $$ \begin{aligned} \|v\|_A &= {(\hspace{-.03in}(} A v,\, v{)\hspace{-.03in})}^{1/2} = (\mathbb{A} \nabla_w v,\, \nabla_w v)^{1/2} \qquad &&\textrm{for all } v\in V_h, \\ \|w\|_{\mathcal{A}} &= ( \mathcal{A} w,\, w)^{1/2} = (\mathbb{A} \nabla w,\, \nabla w)^{1/2} \qquad &&\textrm{for all } w\in V^c_h. \end{aligned} $$ It is well-known that the spectral radius and condition number of operator $\mathcal{A}$ is $O(h^{-2})$~\cite{Bramble93}. We have similar estimate for the operator $A$. Note that the authors of~\cite{LiXie} also give a proof of the order of the condition number. But our proof is different from theirs and seems to be easier. \medskip \begin{lemma} \label{lem:spectralradius} The spectral radius of operator $A$, denoted by $\rho_A=\lambda_{\max}(A)$, and the condition number of operator $A$, denoted by $\kappa(A)$, are both of order $h^{-2}$. \end{lemma} \begin{proof} By the definition of $A$ and Lemma \ref{lem:discreteinverse}, for all $v\in V_h$, $$ {(\hspace{-.03in}(} A v,v{)\hspace{-.03in})} \lesssim \|\nabla_w v\|^2 \lesssim h^{-2} \|v\|_{0,h}^2 = h^{-2} {(\hspace{-.03in}(} v,v{)\hspace{-.03in})}. $$ Because $A$ is symmetric and positive definite with respect to ${(\hspace{-.03in}(} \cdot,\cdot {)\hspace{-.03in})}$, the above inequality implies that $\lambda_{\max}(A)\lesssim h^{-2}$. The discrete Poincar\'e inequality \eqref{eq:discretepoincare1} implies $\lambda_{\min}(A) \gtrsim 1$. Therefore $\kappa(A) = \lambda_{\max}(A)/\lambda_{\min}(A)\lesssim h^{-2}$. To derive a lower bound for $\lambda_{\max}(A)$, we first consider functions in $V_h$ with the form $v = \{0,v_b\}$. In other words, $v_0\equiv 0$. Then, by the definition of discrete norms, Lemma \ref{lem:discretenorm-equivalence} and the fact that $\mathbb{A}$ is uniformly positive definite, for such function $v$ we have $$ \begin{aligned} {(\hspace{-.03in}(} A v,v{)\hspace{-.03in})} &\gtrsim \|\nabla_w v\|^2 \gtrsim |v|_{1,h}^2 = \sum_{K\in \mathcal{T}_h} h^{-1}\|v_b\|_{\partial K}^2 \\ &= h^{-2} \sum_{K\in \mathcal{T}_h} h\|v_b\|_{\partial K}^2= h^{-2} \|v\|_{0,h}^2 = h^{-2} {(\hspace{-.03in}(} v,v{)\hspace{-.03in})}. \end{aligned} $$ Therefore, we must have $\lambda_{\max}(A)\gtrsim h^{-2}$. This implies the spectral radius $\rho_A = \lambda_{\max}(A) = O(h^{-2})$. To get $\lambda_{\min}(A)\lesssim 1$, we chose the eigen-function $w$ of the smallest eigenvalue, $\lambda_1$, of $-\Delta$ with homogeneous Dirichlet boundary condition which satisfies $1=\|\nabla w\| = \sqrt{\lambda_1} \|w\|$. It is well known that $\lambda_1 = O(1)$. We then project $w$ to $V_h$ using the $L^2$-projection, i.e., $w_h = Q_hw$. We estimate the norm of $w_h$ as follows: when $h$ is sufficiently small, by the triangle inequality and Lemma \ref{lem:Qh} one has $$ \|w_h\|\geq \|w\| - \|w-w_h\| \gtrsim \|w\| - Ch\|\nabla w\| = \|w\| - Ch \gtrsim \|w\|, $$ where $C$ is a positive, general constant. By the above inequality and the stability of $Q_h$ in the energy norm, c.f. \eqref{eq:H1stable}, we have $$ \|w_h\|_A \lesssim \|\nabla w\| =\sqrt{\lambda_1} \|w\| \lesssim \|w_h\|. $$ This completes the proof of the lemma. \end{proof} \smallskip \begin{remark} \label{rem:linalgsystem} By the triangle inequality, the trace inequality \eqref{trace} and the inverse inequality, the norm $\|v_h\|_{0,h}$ is equivalent to $\left(\sum_{K\in \mathcal T_h}(\|v_0\|_K^2 + h\|v_b\|_{\partial K}^2)\right)^{1/2}$ in $V_h$. In practice, equation \eqref{eq:wg} can be written as a linear algebraic system by using the canonical bases of $V_h$, i.e. Lagrange bases of $P_j(K)$ and $P_l(e)$ on each $K$ and $e$. Using the standard scaling argument and the equivalent norm of $\|\cdot\|_{0,h}$, it is not hard to see that for any $v_h \in V_h$, one has $\|v_h\|_{0,h}^2 \approx h^{d} \|\underline{v_h}\|_{l^2}^2$, where $\underline{v_h}$ is the vector representation of $v_h$ under the canonical bases and $\|\cdot\|_{l^2}$ is the Euclidean norm of vectors. Then, the stiffness matrix in the linear algebraic system resulting from \eqref{eq:wg}, i.e., the matrix representation of ${(\hspace{-.03in}(} A \cdot,\cdot{)\hspace{-.03in})}$, also has condition number of order $O(h^{-2})$. Thus it is not easy to solve equation \eqref{eq:wg} without efficient preconditioning. \end{remark} \smallskip Next, we introduce the auxiliary space multigrid method for solving equation \eqref{eq:wg}. The idea is to construct a multigrid method using $V_h$ as the ``fine'' space and $V^c_h$ as the ``coarse'' space. Since $\mathcal{A}$ is the discrete Laplacian on the conforming piecewise linear finite element space, the ``coarse'' problem in $V^c_h$ can be solved by many efficient, off-the-shelf solvers such as the standard multigrid solver or a domain decomposition solver. Denote $\mathcal{B}:\: V^c_h\rightarrow V^c_h$ to be such a ``coarse'' solver. It can be either an exact solver or an approximate solver that satisfies certain conditions, which will be given later. Next, on the fine space, we need a ``smoother'' $R:\: V_h\rightarrow V_h$, which is symmetric and positive definite. For example, $R$ can be a Jacobi or symmetric Gauss-Seidel smoother. Finally, to connect the ``coarse'' space with the ``fine'' space, we need a ``prolongation'' operator $\Pi:\:V^c_h\rightarrow V_h$. A ``restriction'' operator $\Pi^t:\: V_h \rightarrow V^c_h$ is consequently defined by $$ ( \Pi^t v, \,w) = {(\hspace{-.03in}(} v,\, \Pi w{)\hspace{-.03in})} \quad\textrm{for } v\in V_h\textrm{ and } w\in V^c_h. $$ Then, the auxiliary space multigrid preconditioner $B:\: V_h\rightarrow V_h$, following the definition in~\cite{xu, Bramble93}, is given by \begin{align} &\textrm{Additive} \qquad &&B = R + \Pi \mathcal{B} \Pi^t, \label{addB}\\ &\textrm{Multiplicative} \qquad && I-BA = (I-RA)(I-\Pi \mathcal{B} \Pi^t)(I-RA). \label{mulB} \end{align} Both the additive and the multiplicative versions define symmetric multigrid solvers/preconditioners. Readers may refer to~\cite{xu92} for the equivalence between symmetric solvers and preconditioners for symmetric problems. Non-symmetric multiplicative multigrid solver can similarly be defined but it cannot be used as a preconditioner. Thus we restrict our attention to the symmetric version. According to~\cite{xu}, the following theorem holds. \medskip \begin{theorem} \label{thm:abscond} Assume that for all $v\in V_h$, $w\in V^c_h$, \begin{align} \rho_A^{-1} {(\hspace{-.03in}(} v,\, v{)\hspace{-.03in})} \lesssim {(\hspace{-.03in}(} R v, \, v{)\hspace{-.03in})} &\lesssim \rho_A^{-1} {(\hspace{-.03in}(} v,\, v {)\hspace{-.03in})}, \label{eq:mg-R}\\ (\mathcal{A} w,\, w) \lesssim (\mathcal{B} \mathcal{A} w,\, \mathcal{A} w) &\lesssim (\mathcal{A} w\, w), \label{eq:mg-cB}\\ \|\Pi w\|_A &\lesssim \|w\|_{\mathcal{A}} \quad\quad\textrm{(stability of $\Pi$)}, \label{eq:mg-Pi} \end{align} and furthermore, assume that there exists a linear operator $P:\, V_h\rightarrow V^c_h$ such that \begin{align} \|P v\|_{\mathcal{A}} &\lesssim \|v\|_A, \quad\qquad\textrm{(stability of $P$)}\label{eq:mg-P1}\\ \|v-\Pi P v\|_{0,h}^2 &\lesssim \rho_A^{-1} \|v\|_A^2 \qquad\textrm{(approximability)}.\label{eq:mg-P2} \end{align} Then the preconditioner $B$ defined in \eqref{addB} or \eqref{mulB} satisfies $$ \kappa (BA) \lesssim O(1). $$ \end{theorem} \begin{remark} Theorem \ref{thm:abscond} states that $B$ is a good preconditioner for $A$ as the condition number of $BA$ is uniformly bounded. We thus can use the preconditioned conjugate gradient (PCG) method with $B$ being an effective preconditioner for solving the linear algebraic equation system associate to $Au = f$. According to~\cite{xu92}, Theorem \ref{thm:abscond} also implies that $I-\omega BA$, where $0<\omega<2/\rho_{BA}$, defines an efficient iterative solver. \end{remark} \medskip Now we shall construct an auxiliary space preconditioner which satisfies all conditions in Theorem \ref{thm:abscond}, namely, inequalities \eqref{eq:mg-R}-\eqref{eq:mg-P2}. It is straight forward to pick $\mathcal{B}$ that satisfies condition \eqref{eq:mg-cB}. For example, $\mathcal{B}$ can be either the direct solver, for which $\mathcal{B}\sim \mathcal{A}^{-1}$, or one step of classical multigrid iteration~\cite{Bramble93} which satisfies condition \eqref{eq:mg-cB}. The smoother $R$ is also easy to define. In view of Remark \ref{rem:linalgsystem}, a Jacobi or a symmetric Gauss-Seidel smoother~\cite{Bramble93} will satisfy condition \eqref{eq:mg-R}. Hence it remains to construct operators $\Pi$ and $P$ that satisfy the conditions \eqref{eq:mg-Pi}-\eqref{eq:mg-P2}. The operator $\Pi$ is actually easy to choose, and we simply define $\Pi=Q_h = \{Q_0,\, Q_b\}$. Note when $V_h$ consists of $W_{j,j}$ elements or $W_{j,j+1}$ elements with $j\ge 1$, it is clear that for all $w\in V^c_h$ and $K\in \mathcal{T}_h$, $(\Pi w)|_K = \{w|_{K_0},\, w|_{\partial K}\}$ which is just the natural inclusion of $\mathcal V_h$ into $V_h$. The stability of $\Pi$ in the energy norm follows immediately from \eqref{eq:H1stable} and the boundedness of the diffusion coefficient $\mathbb A$: \smallskip \begin{lemma} Let $\Pi=Q_h = \{Q_0,\, Q_b\}$. Then $\Pi$ satisfies condition \eqref{eq:mg-Pi}, i.e., $$ \|\Pi w\|_A \lesssim \|w\|_{\mathcal{A}}, \quad \text{ for all } w\in V^c_h. $$ \end{lemma} Next, we construct an operator $P$ that satisfies \eqref{eq:mg-P1} and \eqref{eq:mg-P2}. \begin{definition} Let $0\le \alpha_1,\alpha_2,\ldots, \alpha_k\le 1$ satisfy $\sum_{i=1}^k \alpha_i = 1$, and let $\{c_1,c_2,\ldots, c_k\}$ be a sequence of numbers. The value $\sum_{i=1}^k \alpha_i c_i$ is called a convex combination of $\{c_1,c_2,\ldots, c_k\}$. \end{definition} \medskip A function in $V^c_h$ is completely determined by its value on mesh vertices. Let $v=\{v_0,v_b\} \in V_h$. To define $Pv$, one only needs to specify its value on all mesh vertices. Hence we can define $P$ as follows: on each mesh vertex ${\bf x}$, the value of $Pv ({\bf x})$ is a prescribed convex combination of the values of $v_0({\bf x})$ and $v_b({\bf x})$ on all mesh elements and edges/faces that have ${\bf x}$ as a vertex. Moreover, to preserve the homogeneous boundary condition, when ${\bf x}\in\partial\Omega$, the convex combination shall be constructed such that it only depends on the value of $v_b({\bf x})$ on boundary edges/faces that have ${\bf x}$ as a vertex. Of course, for problems with the homogeneous Dirichlet boundary condition, one can simply set $Pv({\bf x}) = 0$ on boundary vertices. But the current set-up would allow easy extension to non-homogeneous boundary conditions. \medskip \begin{lemma} Operator $P$ satisfies \begin{equation} \label{eq:Pvapproximability} \|v-Pv\|_{0,h}^2 + h^2 |v-Pv|_{1,h}^2 \lesssim h^2 |v|_{1,h}^2,\qquad \textrm{for all } v\in V_h. \end{equation} \end{lemma} \begin{proof} For each $K\in \mathcal{T}_h$, denote by $V(K)$ the vertices of $K$. For each $K\in \mathcal{T}_h$ and $v=\{v_0,v_b\} \in V_h$, denote by $I_{h,K}v_0$ the nodal value interpolation of $v_0$ into $P_1(K)$, i.e., $I_{h,K}v_0\in P_1(K)$ and is identical to $v_0$ on $V(K)$. By the approximation property of nodal value interpolations, the scaling argument, the definition of $P$, the triangle inequality, and the finite overlapping property of quasi-uniform meshes, we have $$ \begin{aligned} &\|v-P v\|_{0,h}^2 \\ =& \sum_{K\in \mathcal{T}_h} \left(|v_0-P v|_{0,K}^2 + h\|v_0-v_b\|_{0,\partial K}^2 \right) \\ \lesssim & \sum_{K\in \mathcal{T}_h} \left(|v_0-I_{h,K} v_0|_{0,K}^2 + \sum_{{\bf x}\in V(K)} h^2 |I_{h,K} v_0 ({\bf x})-Pv({\bf x})|^2 + h\|v_0-v_b\|_{0,\partial K}^2 \right)\\ \lesssim & \sum_{K\in \mathcal{T}_h} \left(h^2|v_0|_{1,K}^2 + \sum_{{\bf x}\in V(K)} h^2 |v_0 ({\bf x})-v_b({\bf x})|^2 + h\|v_0-v_b\|_{0,\partial K}^2 \right)\\ \lesssim & \sum_{K\in \mathcal{T}_h} \left(h^2|v_0|_{1,K}^2 + h \|v_0 - v_b\|_{0,\partial K}^2\right) \\ =& h^2 |v|_{1, h}^2. \end{aligned} $$ Combining the above with the inverse inequality \eqref{eq:discreteinverse} completes the proof of the lemma. \end{proof} \medskip \begin{lemma} \label{lem:Pstability} The operator $P$ satisfies the properties \eqref{eq:mg-P1} and \eqref{eq:mg-P2}. \end{lemma} \begin{proof} By using inequalities \eqref{eq:Pvapproximability} and \eqref{eq:discreteh1semi}, for all $v\in V_h$, we have $$ \|P v\|_{\mathcal{A}}^2 \lesssim |Pv |_{1,h}^2 \lesssim |v-Pv|_{1,h}^2 + |v|_{1,h}^2 \lesssim |v|_{1,h}^2 \lesssim \|v\|_{A}^2. $$ This completes the proof of Inequality \eqref{eq:mg-P1}. We then estimate $\|Pv-\Pi P v\|_{0,h}$. When $j\geq 1$, $\|Pv-\Pi P v\|_{0,h} = 0$ since $\Pi$ is the natural inclusion. We only need to consider the case $j = 0$. Since $\Pi Pv$ is the average of $Pv$, we get $$ \|Pv - \Pi Pv\|_{K}\lesssim h|Pv|_1, \quad \|Pv - \Pi Pv\|_{\partial K}\lesssim h|Pv|_{1,\partial K}, $$ by the average type Poincar\'e inequality. By the scaled trace inequality \eqref{trace} and the fact $|Pv|_{2,K} = 0$ for a piecewise linear function, we can bound $h^{1/2}|Pv|_{1,\partial K} \lesssim |Pv|_{1,K}$. Therefore, we obtain $$ \|Pv-\Pi P v\|_{0,h} \lesssim h|Pv|_{1}= h |Pv|_{1,h}\lesssim h|v|_{1,h}. $$ Then, by the triangle inequality and the coercivity of operator $A$, for all $v\in V_h$, we have $$ \begin{aligned} \|v-\Pi P v\|_{0,h} &\lesssim \|v-Pv\|_{0,h} + \|Pv-\Pi P v\|_{0,h}\lesssim h|v|_{1,h} \lesssim h\|v\|_A. \end{aligned} $$ Combining the above with the estimate $\rho_A = O(h^{-2})$ (see Lemma \ref{lem:spectralradius}), this completes the proof of Inequality \eqref{eq:mg-P2}. \end{proof} \begin{remark} In the proof of Lemma \ref{lem:Pstability}, one may also use Lemma \ref{lem:Qh} and the Poincar\'e inequality to estimate $\|Pv-\Pi P v\|_{0,h}$, i.e., $$ \|Pv-\Pi P v\|_{0,h} \lesssim h\|Pv\|_1 \lesssim h|Pv|_{1}. $$ This requires the Poincar\'e inequality for $Pv$, which is not true for non-homogeneous Dirichlet boundary problems. The current approach avoids such difficulty and can thus be easily extended to non-homogeneous Dirichlet boundary problems or Neumann boundary problems. \end{remark} \smallskip By now, all conditions in Theorem \ref{thm:abscond} have been verified for the given multigrid construction. We summarize it in the following theorem: \begin{theorem} Suppose we have a smoother $R$ and an auxiliary solver $\mathcal B$ satisfying the property: for all $v\in V_h$, $w\in V^c_h$, \begin{align*} \rho_A^{-1} {(\hspace{-.03in}(} v,\, v{)\hspace{-.03in})} \lesssim {(\hspace{-.03in}(} R v, \, v{)\hspace{-.03in})} &\lesssim \rho_A^{-1} {(\hspace{-.03in}(} v,\, v {)\hspace{-.03in})},\\ (\mathcal{A} w,\, w) \lesssim (\mathcal{B} \mathcal{A} w,\, \mathcal{A} w) &\lesssim (\mathcal{A} w\, w). \end{align*} Let $B = R + \Pi \mathcal{B} \Pi^t$ or defined implicitly by the relation $I-BA = (I-RA)(I-\Pi \mathcal{B} \Pi^t)(I-RA)$. Then $B$ is symmetric and positive definite and $\kappa (BA) \lesssim O(1)$. \end{theorem} \medskip \begin{remark} The operator $P$, although its definition seems to be complex, is only needed in the theoretical analysis. In the implementation, one only needs $\mathcal{B}$, $R$ and $\Pi$. It is also well-known that the matrix representation of the restriction operator $\Pi^t$ is just the transpose of the matrix representation of the prolongation operator $\Pi$. \end{remark} \section{Reduced system and its multigrid preconditioner} \label{sec:reduced} By using the Schur complement, the weak Galerkin problem \eqref{eq:wg} can be reduced to a system involving only the degrees of freedom on mesh edges/faces. In this section, we present such a reduced system and construct an auxiliary space multigrid preconditioner for the reduced system. \subsection{Reduced system} Let $$ \begin{aligned} V_0 &= \{v\:|\: v = \{v_0,\,0\}\in V_h\}, \\ V_b &= \{v\:|\: v = \{0,\, v_b\}\in V_h\}, \end{aligned} $$ be two subspaces of $V_h$. Clearly one has $V_h = V_0+V_b$. For any function $v = \{v_0,\,v_b\}\in V_h$, it is convenient to extend the notation of $v_0$ and $v_b$ so that, without ambiguity, $v_0\in V_0$ and $v_b\in V_b$. Functions in $V_0$ and $V_b$ will also often be referred to as $v_0$ and $v_b$, respectively. Then Equation \eqref{eq:wg} can be rewritten into \begin{equation} \label{eq:reduced1} \begin{cases} a_h(u_0, \, v_b) + a_h(u_b, \, v_b) = 0, \quad \qquad \text{ for all } v_b\in V_b,\\ a_h(u_0, \, v_0) + a_h(u_b, \, v_0) = (f,\, v_0)\quad \text{ for all } v_0\in V_0. \end{cases} \end{equation} By choosing a basis of $V_h$, we can obtain a matrix form of \eqref{eq:reduced1}. Let $\mathbf v$ be the vector representation of a weak function $v\in V_h$ and $\mathbf M$ be the matrix representation of an operator $M$ relative to the chosen basis. We can write the matrix form of \eqref{eq:reduced1} as follows \begin{equation}\label{eq:originalmatrix} \begin{pmatrix} \mathbf A_b & \mathbf A_{b0}\\ \mathbf A_{0b} & \mathbf A_{0} \end{pmatrix} \begin{pmatrix} \mathbf u_b\\ \mathbf u_0 \end{pmatrix} = \begin{pmatrix} 0\\ \mathbf f \end{pmatrix}. \end{equation} Note that $\mathbf A_0$ is block-diagonal. We can thus solve $\mathbf u_0$ from the second equation and substitute into the first equation to obtain the Schur complement equation \begin{equation}\label{eq:Schur} (\mathbf A_b - \mathbf A_{b0}\mathbf A_0^{-1}\mathbf A_{0b}) \mathbf u_b = - \mathbf A_0^{-1}\mathbf f. \end{equation} After $\mathbf u_b$ is obtained by solving \eqref{eq:Schur}, the interior part $\mathbf u_0 = \mathbf A_0^{-1}(\mathbf f - \mathbf A_{0b}\mathbf u_b)$ can be computed element-wise. The reduced system \eqref{eq:Schur} involves less degrees of freedom than the original weak Galerkin system \eqref{eq:originalmatrix}. Indeed, the difference between these two degrees of freedom is exactly $\dim(V_0)$, which is equal to $(j+1)(j+2)/2$ times the total number of mesh triangles in two-dimension, and $(j+1)(j+2)(j+3)/6$ times the total number of mesh tetrahedron in three-dimension. More importantly, the Schur complement $\mathbf A_b - \mathbf A_{b0}\mathbf A_0^{-1}\mathbf A_{0b}$ is also a SPD matrix and has the same sparsity as $\mathbf A_b$. Therefore solving the reduced system \eqref{eq:Schur} is more efficient than solving the original system \eqref{eq:originalmatrix} provided a good preconditioner for \eqref{eq:Schur} is available. In the rest of this section, we will construct a fast auxiliary multigrid preconditioner for \eqref{eq:Schur}. Note that the algorithm is implemented in the matrix formulation. The analysis, however, is given in the operator form. In the following we will introduce corresponding operators. We first introduce an $a_h(\cdot, \cdot)$-orthogonal projector $P_0$ from $V_b$ to $V_0$ as follows: For $v_b\in V_b$, define $P_0v_b\in V_0$ such that $$ a_h(P_0 v_b,\, \zeta_0 ) = a_h(v_b,\, \zeta_0 ) \qquad\textrm{for all } \zeta_0\in V_0. $$ It is not hard to see that $\|(I-P_0)v_b\|_{0,h} = \|\{-P_0v_b,v_b\}\|_{0,h}$ is a well-defined norm on $V_b$. In the following analysis we shall always equip $V_b$ with this new norm and $V_0$ with the inherited norm $\|\cdot\|_{0,h}$. By the trace inequality, the inverse inequality and the definition of $\|\cdot\|_{0,h}$, one has $$ \|P_0v_b\|_{0,h} \lesssim \|P_0 v_b\| \lesssim \|\{-P_0v_b,v_b\}\|_{0,h} = \|(I-P_0)v_b\|_{0,h}, $$ which implies that $P_0:\: V_b\rightarrow V_0$ is a bounded linear operator under the newly assigned norms. Denote by $V_0'$ and $V_b'$ the space of bounded linear functionals on $V_0$ and $V_b$, respectively. Then the bounded linear operator $P_0$ induces a bounded dual operator $P_0': V_0'\to V_b'$, i.e., for $F\in V_0'$, $\langle P_0'F, v_b \rangle \triangleq \langle F, P_0 v_b\rangle $ for all $v_b \in V_b$. In particular, let $F$ be defined by $\langle F, \cdot\rangle = (f, \cdot)$ for $f\in L^2(\Omega)$, then one has $\langle P_0'F, v_b \rangle = (f, P_0v_b)$. We claim, and will prove later, that the operator form of the Schur complement equation \eqref{eq:Schur} is \begin{equation} \label{eq:reduced} \begin{aligned} a_h((I - P_0) u_b,\, v_b) = -\langle P_0'F, \, v_b\rangle, \quad \text{for all } v_b\in V_b. \end{aligned} \end{equation} Note that by the property of the projection $P_0$, Equation \eqref{eq:reduced} can also be written into the symmetric form $a_h((I - P_0) u_b,\, (I - P_0)v_b) = -\langle P_0'F, v_b\rangle$ for all $v_b\in V_b$. To prove this, we first define a linear operator $A_0^{-1}: L^2(\Omega)\rightarrow V_0$ by: for a function $g \in L^2(\Omega)$, one has $A_0^{-1} g\in V_0$ such that $$ a_h(A_0^{-1} g, \, v_0) = (g,\, v_0)\qquad \textrm{for all }v_0\in V_0. $$ The well-posedness of $A_0^{-1}$ follows directly from the coercivity of $a_h(\cdot,\cdot)$ on $V_h$, and consequently on its subspace $V_0$. Moreover, the restriction of $A_0^{-1}$ to $V_0$ is symmetric and positive definite. Noticing that $\nabla_w v_0$ is locally defined on each mesh element, it is clear that $A_0^{-1}$ is also locally defined on each mesh element. Denote by $\nabla_h\cdot$ the piecewise divergence operator on $\Sigma_h$, and by $\mathbb Q_h:\: L^2(\Omega)^d \rightarrow \Sigma_h$ the $L^2$ projection. Using the above notation and the definition of $\nabla_w$ , the second equation in \eqref{eq:reduced1} implies that \begin{equation} \label{eq:u0} \begin{aligned} (\mathbb{A} \nabla_w u_0,\,\nabla_w v_0) &= (f,\, v_0) - (\mathbb{A} \nabla_w u_b, \, \nabla_w v_0) \\ &= (f,\, v_0) + (\nabla_h\cdot (\mathbb Q_h\mathbb{A} \nabla_w u_b),\, v_0), \end{aligned} \end{equation} which leads to \begin{equation} \label{eq:u0-2} u_0 = A_0^{-1} (f + \nabla_h\cdot (\mathbb Q_h\mathbb{A} \nabla_w u_b) ). \end{equation} \medskip Next, we note that the projection $P_0$ is identical to $-A_0^{-1} \nabla_h\cdot (\mathbb Q_h\mathbb{A} \nabla_w)$ on $V_b$: \begin{lemma} \label{lem:orthogonal} The orthogonal operator $P_0:\:V_b\rightarrow V_0$ $$ P_0 v_b = -A_0^{-1} \nabla_h\cdot (\mathbb Q_h\mathbb{A} \nabla_w v_b)\qquad \textrm{for all } v_b\in V_b. $$ \end{lemma} \begin{proof} By the definition of weak gradient $\nabla_w$ and $A_0^{-1}$, we have \begin{align*} (\mathbb{A} \nabla_w v_b,\, \nabla_w \zeta_0 ) &= - (\nabla_h\cdot (\mathbb Q_h\mathbb{A} \nabla_w v_b),\, \zeta_0 ) \\ &= -(\mathbb{A}\nabla_w A_0^{-1} \nabla_h\cdot (\mathbb Q_h\mathbb{A} \nabla_w v_b),\, \nabla_w \zeta_0 ). \end{align*} By the definition of $P_0$, we then complete the proof of the lemma. \end{proof} \smallskip \begin{remark} The operator $P_0$ corresponds to the matrix $\mathbf A_0^{-1}\mathbf A_{0b}$. \end{remark} \smallskip Now, by \eqref{eq:u0-2} and Lemma \ref{lem:orthogonal}, one has $u_0=A_0^{-1} f - P_0 u_b$. Substituting this into the first equation of \eqref{eq:reduced1} gives \begin{align*} a_h((I-P_0)u_b,\, v_b) &= -a_h(A_0^{-1}f, \,v_b) = -a_h(A_0^{-1}f, \,P_0 v_b) \\ & = -(f, \, P_0 v_b) = -\langle P_0'F, v_b \rangle. \end{align*} This completes the derivation of the reduced problem \eqref{eq:reduced} from the original problem \eqref{eq:wg}. Here we emphasize again that $P_0'F \in V_b'$ is bounded in the sense that \begin{equation} \label{eq:boundedlf} |\langle P_0'F, v_b \rangle| \lesssim \|(I-P_0)v_b\|_{0,h}. \end{equation} We will further reformulate the reduced system \eqref{eq:reduced}. To this end, we define a subspace of $V_h$ as $V_r = \{ v_r\, | \, v_r = (I-P_0)v_b = \{-P_0 v_b, v_b\}\textrm{ for all } v_b \in V_b\}$, which is just the graph of $V_b$ under $I - P_0$. The space $V_r$ inherits the norm $\|\cdot\|_{0,h}$ from $V_h$, and hence $V_r$ and $V_b$ (equipped with the norm $\|(I-P_0)\cdot\|_{0,h}$) are clearly isomorphic under the mapping $I-P_0:\: V_b\rightarrow V_r$. Moreover, the right-hand side of Equation \eqref{eq:reduced} can be written into $$ \begin{aligned} -\langle P_0'F, \, v_b\rangle &= -\langle \{0,P_0'F\},\, \{-P_0 v_b,v_b\}\rangle = -\langle \{0,P_0'F\},\, v_r \rangle \\ &\triangleq \langle \mathcal{F}, \, v_r \rangle, \end{aligned} $$ where $\mathcal{F}$ is a bounded linear functional on $V_r$ according to \eqref{eq:boundedlf}. By using Lemma \ref{lem:orthogonal} and combining the above analysis, Equation \eqref{eq:reduced} can now be rewritten into: Find $u_r\in V_r$ such that \begin{equation}\label{eq:reducedeq} a_h(u_r,\, v_r) = \langle \mathcal{F}, v_r\rangle, \quad \text{for all }v_r\in V_r. \end{equation} The well-posedness of \eqref{eq:reducedeq} then follows from the continuity and coercivity of the bilinear form $a_h(\cdot, \cdot)$ restricted to $V_r$ and the fact that $\mathcal{F}$ is a bounded linear functional on $V_r$. \subsection{Auxiliary space preconditioner for the reduced system} Now we are able to consider an auxiliary space multigrid preconditioner for the reduced system \eqref{eq:reducedeq}, using again the $H^1$ conforming piecewise linear finite element space as the auxiliary space. Denote by $A_r$ the restriction of operator $A$, defined in \eqref{eq:Adefinition}, to the subspace $V_r$. That is, $A_r:\: V_r\rightarrow V_r$ is defined by $$ {(\hspace{-.03in}(} A_r u,\, v{)\hspace{-.03in})} = a_h(u,\, v)\quad\textrm{for all }v\in V_r. $$ To apply Theorem \ref{thm:abscond}, we define a prolongation operator $\Pi_r:\: V^c_h\rightarrow V_r$ and a linear operator $P_r:\: V_r\rightarrowV^c_h$ by $$ \Pi_r = (I-P_0)Q_b \quad\textrm{and}\quad P_r = P|_{V_r}. $$ \begin{lemma} Both $\Pi_r$ and $P_r$ are stable in the energy norm, i.e., \begin{align} \label{Pistable}\| \Pi_r v \|_{A} \lesssim \|v\|_{\mathcal{A}},\qquad\textrm{for all }v\inV^c_h\\ \label{Pstable} \|P_r v_r\|_{\mathcal{A}} \lesssim \|v_r\|_A,\qquad\textrm{for all }v_r\in V_r. \end{align} \end{lemma} \begin{proof} The stability of $\Pi_r$ follows from the property of $P_0$ and the stability \eqref{eq:H1stable} of $Q_h$: $$ \begin{aligned} \| \Pi_r v \|_{A}^2 &= \|(I-P_0)Q_b v\|_A^2 = (\mathbb{A} \nabla_w (I-P_0)Q_b v,\, \nabla_w (Q_bv +Q_0 v)) \\ &\lesssim \| \Pi_r v \|_{A} \|Q_h v\|_A \lesssim \| \Pi_r v \|_{A} \|v\|_{\mathcal{A}}. \end{aligned} $$ The stability of $P_r$ simply follows from that of $P$. This completes the proof of the lemma. \end{proof} To verify the approximation property, we first explore the relation between $Q_hw$ and $\Pi_r w$ for $w\in \mathcal V_h$. It turns out that $Q_hw = \Pi_r w$ for all $w\in \mathcal V_h$ when the diffusion coefficient matrix $\mathbb A$ is piecewise constant. \begin{lemma} When $\mathbb A$ is piecewise constant, we have for all $w\in \mathcal V_h$, $$ Q_hw = \Pi_r w. $$ \end{lemma} \begin{proof} Recall that $\Pi_r w = (I-P_0) Q_b w$. Since $P_0$ is the orthogonal projection, we have $$ (\mathbb{A}\nabla_w\Pi_r w ,\, \nabla_w \zeta_0) = (\mathbb{A}\nabla_w (I-P_0) Q_b w),\, \nabla_w \zeta_0) = 0\qquad\textrm{for all }\zeta_0\in V_0. $$ On the other hand, using the relation \eqref{eq:QDDQ} and the fact that both $\nabla w$ and $\mathbb A$ are piecewise constant, \begin{align*} (\mathbb{A} \nabla_w Q_h w, \, \nabla_w \zeta_0)_K = (\mathbb{A}\mathbb Q_h \nabla w, \, \nabla_w \zeta_0)_K = (\mathbb{A}\nabla w, \, \nabla_w \zeta_0)_K\\ = - (\nabla_h\cdot(\mathbb{A} \nabla w),\, \zeta_0)_K = 0, \qquad\textrm{for all }\zeta_0\in V_0. \end{align*} Therefore \begin{equation}\label{orth} a_h(\Pi_r w - Q_h w, \zeta_0) = 0, \quad \text{for all } \zeta_0\in V_0 \end{equation} The fact $\Pi_r w - Q_h w \in V_0$ and the orthogonality~\eqref{orth} implies $\Pi_r w= Q_h w$. \end{proof} \medskip Similar to the analysis in Section \ref{sec:multigrid}, we can establish the following results. \begin{lemma} \label{lem:reducedspectralradius} Suppose $\mathbb{A}$ is piecewise constant and the space $V_r$ is non-trivial, i.e., the triangulation contains at least one interior vertex. Then the spectral radius of operator $A_r$, denoted by $\rho_{A_r}$, is of order $h^{-2}$. \end{lemma} \begin{proof} Recall that $$\rho_{A_r} = \lambda_{\max}(A_r) = \max_{v\in V_r}\frac{{(\hspace{-.03in}(} A_r v, v{)\hspace{-.03in})}}{{(\hspace{-.03in}(} v, v{)\hspace{-.03in})}} = \max_{v\in V_r}\frac{{(\hspace{-.03in}(} Av, v{)\hspace{-.03in})} }{{(\hspace{-.03in}(} v, v{)\hspace{-.03in})}}.$$ Since $V_r\subset V_h$, we immediately get $\rho_{A_r}\leq \rho_A\lesssim h^{-2}$. To show the lower bound, we pick a hat function $w\in \mathcal V_h$. By the standard scaling argument, \begin{equation}\label{eq:wscaling} \|w\|\lesssim h|\nabla w|. \end{equation} We then chose $v = \Pi_r w \in V_r$ and estimate its norms. First \begin{equation}\label{eq:vl2} \|v\|_{0,h} = \|\Pi_r w\|_{0,h} = \|Q_h w\|_{0,h}\lesssim \|w\|. \end{equation} Second, as $\nabla w$ is piecewise constant, $\mathbb Q_h \nabla w = \nabla w$ and \begin{equation}\label{eq:vh1} \|\nabla w\| = \|\mathbb Q_h \nabla w\| = \|\nabla _w Q_h w\| = \|\nabla _w \Pi_r w\| \lesssim (A_r v, v) ^{1/2}. \end{equation} Combining \eqref{eq:vl2}, \eqref{eq:vh1}, and \eqref{eq:wscaling}, we obtain $$ h^{-2}\|v\|_{0,h}^2 \lesssim (A_r v, v), $$ which implies $\rho_{A_r}\gtrsim h^{-2}$. \end{proof} Now we are able to derive the following approximation property: \begin{lemma} Under the same assumptions as in Lemma \ref{lem:reducedspectralradius}, one has $$ \|v_r - \Pi_r P_r v_r\|_{0,h}\lesssim \rho_{A_r}^{-1/2}\|v_r\|_A \qquad \textrm{for all }v_r\in V_r. $$ \end{lemma} \begin{proof} By the triangular inequality and Equation \eqref{eq:Pvapproximability}, one has $$ \begin{aligned} \|v_r - \Pi_r P_r v_r\|_{0,h} &= \|(I-P_0)v_b - \Pi_r P (I-P_0) v_b\|_{0,h} \\[2mm] &\lesssim \|(I-P_0)v_b - P(I-P_0)v_b\|_{0,h} + \|w - Q_h w\|_{0,h}\\ &\lesssim h |(I-P_0)v_b|_{1,h} + h\|w\|_1,\\ &\lesssim h |(I-P_0)v_b|_{1,h}, \end{aligned} $$ where we conveniently denote $w = P(I-P_0)v_b \in V^c_h$ and use $\Pi_r w = Q_hw $. In the last step, we have used $$ h \|w\|_1 \lesssim h|w|_1= h|P(I-P_0)v_b|_{1,h} \lesssim h|(I-P_0)v_b|_{1,h}. $$ Combining the above and using Lemma \ref{lem:discretenorm-equivalence} give $$ \|v_r - \Pi_r P_r v_r\|_{0,h} \lesssim h |(I-P_0)v_b|_{1,h} \lesssim h \|v_r\|_{A}. $$ According to Lemma \ref{lem:reducedspectralradius}, $\rho_{A_r}=O(h^{-2})$. This completes the proof of the lemma. \end{proof} Finally, for variable efficient $\mathbb A$, if we denoted by $\bar{\mathbb A}$, the piecewise constant approximation of $\mathbb A$, then it is easy to show $$ (\mathbb A \nabla_w v, \nabla_w v) \lesssim (\bar{\mathbb A}\nabla_w v, \nabla_w v) \lesssim (\mathbb A \nabla_w v, \nabla_w v), \quad \text{for all }v \in V_h. $$ Therefore, a good preconditioner for the piecewise constant case will lead to a good preconditioner for the variable case. We are able to claim that, the auxiliary space multigrid preconditioner for the reduced system \eqref{eq:reducedeq} again yields a preconditioner system with condition number of $O(1)$. \begin{theorem} Suppose we have a smoother $R$ and auxiliary solver $\mathcal B$ satisfying the property: for all $v\in V_r$, $w\in V^c_h$, \begin{align*} \rho_{A_r}^{-1} {(\hspace{-.03in}(} v,\, v{)\hspace{-.03in})} \lesssim {(\hspace{-.03in}(} R v, \, v{)\hspace{-.03in})} &\lesssim \rho_{A_r}^{-1} {(\hspace{-.03in}(} v,\, v {)\hspace{-.03in})},\\ (\mathcal{A} w,\, w) \lesssim (\mathcal{B} \mathcal{A} w,\, \mathcal{A} w) &\lesssim (\mathcal{A} w\, w). \end{align*} Let $B = R + \Pi \mathcal{B} \Pi^t$ or defined implicitly by the relation $I-BA_r = (I-RA_r)(I-\Pi \mathcal{B} \Pi^t)(I-RA_r)$. Then $B$ is symmetric and positive definite and $\kappa (BA_r) \lesssim O(1)$. \end{theorem} \section{Numerical results} \label{sec:numerical} In this section, we examine the effectiveness of the auxiliary space multigrid preconditioner using several numerical examples. In all numerical experiments, we use the symmetric Gauss-Seidel smoothers as $R$ and the multiplicative version of the multigrid preconditioner. It is known that the multiplicative version multigrid usually performs better than the corresponding additive version. The simulation is implemented using the MATLAB software package $i$FEM~\cite{Chen.L2008c}. The matrix $A$ for the lowest order weak Galerkin discretization, i.e., $P_0-P_0$ element, is assembled and the matrix $\mathcal A$ for the auxiliary problem using $P1$ element is obtained through the triple product $\mathcal A = \Pi^tA\Pi$ where $\Pi: \mathcal V_h \to V_h$ is the simple average. By doing so, there is no need to repeat the assembling procedure to get $\mathcal A$ and the implementation is more algebraic. After that, the matrices in coarse levels are obtained by the triple product using the standard prolongation and restriction operators of linear elements on hierarchical meshes. We use PCG with the auxiliary space multigrid preconditioner. The stopping criteria for all iterations are reached when the relative error of the residual is less than $10^{-8}$. We report results for the original system and the reduced system, respectively. Since the main purpose of these numerical results is to examine the efficiency of the auxiliary space preconditioner instead of testing the accuracy of the weak Galerkin approximation, in the report we omit the approximation error part. Because of this, there is no need to list the exact solution for each test problem. \subsection*{Example 1} We first consider the Poisson equation defined on a circular mesh of the unit disk. The coarsest mesh is shown in Fig.~\ref{fig:2Dmesh} (a). A sequence of meshes are obtained by several uniformly regular refinements, i.e., a triangle is divided into four congruent four triangles by connecting middle points of edges, of the coarsest mesh. Results are summarized in Table \ref{example1}. \subsection*{Example 2} Next, we consider a variable coefficient problem with an oscillating coefficient: $$ -\nabla\cdot \left( 2(2+\sin(10\pi x)\sin(10\pi y)) \nabla u\right) = f $$ on $[0,1]\times[0,1]$. The coarsest mesh has size $h=1/4$ and is shown in Fig.~\ref{fig:2Dmesh} (b). Fourth order quadrature is used when assembling the stiffness matrix. Results are summarized in Table \ref{example2}. \subsection*{Example 3} We consider a test problem on an $L$-shaped domain obtained by subtracting $[0,1]\times[-1,0]$ from $(-1,1)\times(-1,1)$. The Poisson equation on such a domain has $H^{3/2}$-regularity. Adaptive finite element method based on {\it a posteriori} error estimator constructed in~\cite{Chen2013} is used. A sample adaptive mesh obtained by bisection refinement is shown in Fig.~\ref{fig:2Dmesh} (c). For bisection grids, we apply coarsening algorithm developed in~\cite{Chen2010a} to obtain a hierarchy of meshes. In Table \ref{example3}, only results on some selected adaptive meshes are reported since the full list of adaptive meshes are long and the performance remains similar for all meshes. \begin{figure}[htbp] \label{fig:2Dmesh} \subfigure[Initial grid of Example 1]{ \begin{minipage}[t]{0.33\linewidth} \centering \includegraphics*[width=3.1cm]{circlemesh.pdf} \end{minipage} \subfigure[Initial grid of Example 2] {\begin{minipage}[t]{0.32\linewidth} \centering \includegraphics*[width=3cm]{squaremesh.pdf} \end{minipage} \subfigure[An adaptive grid of Example 3] {\begin{minipage}[t]{0.34\linewidth} \centering \includegraphics*[width=3cm]{Lshapemesh.pdf} \end{minipage}} \caption{Meshes in Example 1, 2, 3.} \end{figure} \begin{table}[htdp] \caption{Iteration steps and CPU time (in seconds) for Example 1. The left table is for the original system and the right table is for the reduced system.} \begin{center} \begin{tabular}{c c c } \hline \hline Dof & Steps & Time\\ \hline 3446 & 13 & 0.052\\ \hline 13692 & 13 & 0.11\\ \hline 54584 & 13 & 0.41\\ \hline 217968 & 13 & 1.8\\ \hline 871136 & 13 & 8\\ \hline \hline \end{tabular} \quad \quad \begin{tabular}{c c c } \hline \hline Dof & Steps & Time\\ \hline 2086 & 8 & 0.02 \\ \hline 8252 & 8 & 0.053 \\ \hline 32824 & 8 & 0.17 \\ \hline 130928 & 8 & 0.66 \\ \hline 522976 & 8 & 2.9 \\ \hline \hline \end{tabular} \end{center} \label{example1} \end{table} \begin{table}[htdp] \caption{Iteration steps and CPU time (in seconds) for Example 2. The left table is for the original system and the right table is for the reduced system.} \begin{center} \begin{tabular}{c c c } \hline \hline Dof & Steps & Time\\ \hline 1312 & 13 & 0.017 \\ \hline 5184 & 13 & 0.048 \\ \hline 20608 & 14 & 0.17 \\ \hline 82176 & 14 & 0.63 \\ \hline 328192 & 14 & 2.7 \\ \hline \hline \end{tabular} \quad \quad \begin{tabular}{c c c } \hline \hline Dof & Steps & Time\\ \hline 800 & 9 & 0.013 \\ \hline 3136 & 9 & 0.036 \\ \hline 12416 & 10 & 0.082 \\ \hline 49408 & 9 & 0.27 \\ \hline 197120 & 9 & 1.1 \\ \hline \hline \end{tabular} \end{center} \label{example2} \end{table}% \begin{table}[htdp] \caption{Iteration steps and CPU time (in seconds) for Example 3. The left table is for the original system and the right table is for the reduced system.} \begin{center} \begin{tabular}{c c c } \hline \hline Dof & Steps & Time\\ \hline 256 & 13 & 0.0099 \\ \hline 574 & 13 & 0.023 \\ \hline 1091 & 13 & 0.035\\ \hline 2177 & 13 & 0.058\\ \hline 4398 & 13 & 0.11\\ \hline 8642 & 13 & 0.16\\ \hline 10742 & 13 & 0.2\\ \hline \hline \end{tabular} \quad \quad \begin{tabular}{c c c } \hline \hline Dof & Steps & Time\\ \hline 160 & 8 & 0.0088\\ \hline 352 & 9 & 0.014\\ \hline 663 & 10 & 0.023\\ \hline 1317 & 10 & 0.036\\ \hline 2656 & 10 & 0.067\\ \hline 5206 & 9 & 0.091\\ \hline 6470 & 8 & 0.094\\ \hline \hline \end{tabular} \end{center} \label{example3} \end{table}% \begin{figure}[htbp] \label{fig:3Dmesh} \subfigure[Initial grid of Example 4]{ \begin{minipage}[t]{0.45\linewidth} \centering \includegraphics*[width=4.2cm]{cubemesh.pdf} \end{minipage} \subfigure[Domain of Example 5] {\begin{minipage}[t]{0.45\linewidth} \centering \includegraphics*[width=4.1cm]{jump3d.pdf} \end{minipage}} \caption{The coefficients $a_{1} =a_{2}=1$ in the gray domains $\Omega _1$ and $\Omega _2,$ and $a_{3} = \varepsilon$ in the rest of the domain.} \end{figure} \subsection*{Example 4} We consider the Poisson equation defined on the cube $\Omega = (-1,1)^3$. The coarsest mesh is shown in Fig.~\ref{fig:3Dmesh} (a). A sequence of meshes are obtained by several uniformly regular refinements, i.e., a tetrahedron is divided into $8$ small tetrahedron by connecting middle points of edges, of the coarsest mesh. Results are summarized in Table \ref{example4}. \subsection*{Example 5} We consider the elliptic equation with jump coefficients~\cite{Xu.J1991,Xu.J;Zhu.Y2008}. Let $\Omega = (-1,1)^3$ and the diffusion coefficient $a(x)I$ be defined such that $a(x)$ is equal to the constants $a_1=a_2=1$ and $a_3=\varepsilon$ on the three regions $\Omega_1,\;\Omega_2$ and $\Omega_3$ respectively (see Figure~\ref{fig:3Dmesh} (b)), where $$ \Omega_1=(-0.5, 0)^3,\Omega_2=(0,0.5)^3 \; \hbox{ and }\; \Omega_3=\Omega\setminus (\overline{\Omega}_1\cup\overline{\Omega}_2). $$ A sequence of meshes are obtained by several uniformly regular refinements of the coarsest mesh. We choose $f=1$ and impose the following boundary conditions: Dirichlet conditions $$u_{\{-1\}\times[-1,1]\times[-1,1]}=0, \quad u_{\{1\}\times[-1,1]\times[-1,1]} =1,$$ and homogeneous Neumann boundary conditions on the remaining boundary. We test the robustness of our solver as the coefficient $\epsilon$ changes. Only the reduced system is solved in this example. Results are summarized in Table \ref{example5}. \begin{table}[htdp] \caption{Iteration steps and CPU time (in seconds) for Example 4. The left table is for the original system and the right table is for the reduced system.} \begin{center} \begin{tabular}{c c c } \hline \hline Dof & Steps & Time\\ \hline 1248 & 16 & 0.018\\ \hline 9600 & 18 & 0.1\\ \hline 75264 & 18 & 1\\ \hline 595968 & 19 & 10\\ \hline 4743168 & 19 & 100\\ \hline \hline \end{tabular} \quad \quad \begin{tabular}{c c c } \hline \hline Dof & Steps & Time\\ \hline 864 & 11 & 0.058\\ \hline 6528 & 12 & 0.054\\ \hline 50688 & 13 & 0.41\\ \hline 399360 & 13 & 4\\ \hline 3170304 & 13 & 36\\ \hline \hline \end{tabular} \end{center} \label{example4} \end{table}% \begin{table} \caption{Iteration steps for Example 5. Only results for solving the reduced system is presented.} \begin{center} \begin{tabular}{c c c c c c} \hline \hline Dof & $\epsilon = 10^{-4}$ & $\epsilon = 10^{-2}$ & $\epsilon = 1$ & $\epsilon = 10^{2}$ & $\epsilon = 10^{4}$\\ \hline 864 & 36 & 22 & 13 & 13 & 13\\ \hline 6528 & 33 & 21 & 13 & 13 & 13\\ \hline 50688 & 32 & 21 & 13 & 13 & 13\\ \hline 399360 & 34 & 21 & 13 & 13 & 13\\ \hline 3170304 & 34 & 21 & 13 & 13 & 13\\ \hline \hline \end{tabular} \end{center} \label{example5} \end{table} From these experiments we may draw the following conclusions: \begin{enumerate} \item In all examples, the auxiliary space preconditioner works well for the linear system arising from discretization of the lowest order weak Galerkin method. The fluctuation of iteration steps of the PCG method applied to systems with different sizes is small which implies the condition number of the preconditioned system is uniformly bounded. \item The solver for the reduced system is more efficient than the original system. The size of the reduced system is around two thirds of the original one and the time for solving the reduced system is around half of the original one. This shows the efficiency gained by working on the reduced system. \item Although our theory is developed for quasi-uniform meshes, the third example indicates that our solver works well for adaptive grids and elliptic equations with less regularity. \end{enumerate}
\section{Introduction} \label{sec: introduction} Let $S$ be a connected surface and $\xi \in \mathbb{Z}_{\geq 1}$. The \emph{configuration space of $\xi$ points in $S$} is the space of subsets of $S$ of size exactly $\xi$. A \emph{surface braid group} is a fundamental group, $\mathrm{Br}_\xi(S):= \pi_1(\mathrm{Conf}_{\xi}(S))$. When $S = D^2$ is a disk, then this is the usual Artin braid group with $\xi$ strands. In this paper, we are interested in studying particular subgroups of surface braid groups which we define as follows. The \emph{$n$th ordered $\xi$-configuration space of $S$} is \[ \mathrm{PConf}_n^\xi(S) := \{ ( \bm{p}_1, \ldots, \bm{p}_n ) \subset \mathrm{Conf}_\xi(S)^{\times n} \; \vert \; \bm{p}_i \cap \bm{p}_j = \emptyset \mbox{ for } i \neq j \}. \] The symmetric group $\Sigma_n$ acts on $\mathrm{PConf}_n^\xi(S)$ by: if $\sigma \in \Sigma_n$ then $\sigma.(\bm{p}_1, \ldots, \bm{p}_n) = (\bm{p}_{\sigma(1)}, \ldots, \bm{p}_{\sigma(n)})$. \begin{definition} The \emph{$\xi$-configuration space} of $S$ is the quotient \[ \mathrm{Conf}_n^\xi(S) := \frac{\mathrm{PConf}_n^\xi(S)}{\Sigma_n}. \] \end{definition} One way to view this space is that it is the usual configuration space, $\mathrm{Conf}_{n\xi}(S)$, of $n\xi$ points in $S$, where the points have been partitioned into $n$ subsets of size $\xi$. There is a covering map \[ \mathrm{Conf}_n^\xi(S) \to \mathrm{Conf}_{n\xi}(S) \] that forgets the partition of points into subsets. This is a finite sheeted covering map whose fibres correspond to the number of ways to group $n \xi$ points into subsets of size $\xi$. In our pictures, we will represent points being in different subsets by using different shapes. It is also natural to think of points being ``coloured". \begin{definition} The \emph{$\xi$-surface braid group} of $S$ is the fundamental group \[ \mathrm{Br}^\xi_n(S) := \pi_1(\mathrm{Conf}_n^\xi(S)). \] \end{definition} In \cite{fn62}, Fadell and Neuwirth show that $\mathrm{Conf}_{n\xi}(S)$ is $K(\pi,1)$. The long exact sequence in homotopy groups then shows that $\mathrm{Conf}_n^\xi(S)$ is also $K(\pi,1)$. There is therefore an isomorphism \[ H_*(\mathrm{Conf}_n^\xi(S) ; \mathbb{Z}) \cong H_*(\mathrm{Br}_n^\xi(S) ; \mathbb{Z}) \] between singular homology and group homology. The upshot of this isomorphism is that studying the homology of $\mathrm{Conf}_n^\xi(S)$ is the same as studying the homology of $\mathrm{Br}_n^\xi(S)$. We will freely switch between the two. The goal of this paper is to prove the following two theorems. \begin{reptheorem}{thm: surface braid stability} Let $S$ be the interior of a surface with boundary. There is a map $\mathrm{stab} : \mathrm{Br}^\xi_n(S) \rightarrow \mathrm{Br}^\xi_{n+1}(S)$ such that the induced map \[ \mathrm{stab}_*: H_k(\mathrm{Br}^\xi_n(S) ; \mathbb{Z}) \rightarrow H_k(\mathrm{Br}^\xi_{n+1}(S) ; \mathbb{Z}) \] is an isomorphism for $2k \leq n$. \end{reptheorem} \begin{reptheorem}{thm: surface braid closed} Let $S$ be any surface (open or closed). \[ H_*(\mathrm{Br}^\xi_n(S) ; \mathbb{Q}) \cong H_*(\mathrm{Br}^\xi_{n+1}(S) ; \mathbb{Q}) \] for $2k \leq n$. The isomorphism can be realised as a transfer map \[t_{n+1} : H_k(\mathrm{Br}_{n+1}^\xi (S); \mathbb{Q}) \to H_k(\mathrm{Br}_n^\xi(S);\mathbb{Q})\] which we define in \cref{sec: braid stability closed}. \end{reptheorem} The main technical result that we will use is a high connectivity result for certain simplicial complexes which we call \emph{fern complexes} and are defined in \cref{sec: fern complex}. The groups $\mathrm{Br}_n^\xi(S)$ will act on these fern complexes and from there, the literature has a well oiled machine for proving homological stability. In \cref{sec: open}, we will make use of an axiomatisation of this machine by Hatcher and Wahl, which is Theorem 5.1 of \cite{hw10} to prove \cref{thm: surface braid stability}. \subsection{A brief history} Homological stability for braid groups was first proven by Arnol'd in \cite{arnold69}. Homological stability was then extended to configuration spaces of open manifolds by the work of Segal and McDuff in the 70's \cite{segal73, mcduff75, segal79}. Randal-Williams and Church have recently shown that configuration spaces of closed manifolds satisfy homological stability with rational coefficients \cite{orw13, church12}. This paper studies the homology of natural covering space of configuration spaces. Thus the main theorems can be regarded as a homological stability theorem for braid groups with certain twisted coefficients. Configuration spaces for manifolds other than points have also been studied in the literature. For example, the configuration space of circles in $\mathbb{R}^3$ has been well studied (see for example \cite{bh13}). Its fundamental group is sometimes called the symmetric automorphism group, $\Sigma \mathrm{Aut} F_n$, or just the ring group $R_n$. Homological stability for $R_n$ was proven by Hatcher and Wahl in \cite{hw10}. Homological stability for $\xi$-configuration spaces gives an example of homological stability for manifolds where the manifold being stabilised by is disconnected. In his doctoral thesis, Palmer proves homological stability for configurations of (possibly disconnected) manifolds in a larger ambient manifold under certain dimension conditions \cite{palmer13b}. However, the dimesnion conditions do not apply to the case of $0$-manifolds in a surface so that this paper serves also to fill this curious gap in the literature. \subsection{Outline} In \cref{sec: fern complex} we define a simplicial complex on which $\mathrm{Br}_n^\xi(S)$ will act and show that this simplicial complex is highly connected. In \cref{sec: stab} we give a definition of the stabilisation map for $\xi$-configuration spaces, which can be appropriately identified with the stabilisation map for $\xi$-braids. In \cref{sec: open} we prove \cref{thm: surface braid stability}. In \cref{sec: braid stability closed} we define the transfer map for $\xi$-configuration spaces and use it to prove \cref{thm: surface braid closed}. \subsection{Acknowledgements} I would like to thank Nathalie Wahl and Craig Westerland for many useful discussions relating to this work. I would also like to thank Jeffrey Bailes for reading earlier drafts of this paper. This work formed part of my work towards a Ph.D at the University of Melbourne. It was a wonderful place to do research. \section{Fern complexes} \label{sec: fern complex} In this section we will define the simplicial complexes that our $\xi$-braids will act on and show that they are highly connected. We will make use of some of the theory of simplicial complexes which is well summarised in the appendix of \cite{wahl12}. Let $\xi \in \mathbb{Z}_{\geq 1}$ be fixed. Let $S$ be a surface with boundary with $n\xi$ marked points in its interior. Further let $\Delta_1, \ldots, \Delta_n$ be a partition of those marked points into disjoint sets of size $\xi$. Lastly, fix a base point $*$ in the boundary $\partial S$. By an \emph{arc}, we will mean an embedded path in $S$ with one endpoint meeting the boundary of $D$ at $*$ transversally and the other at one of the marked points in $S$. Simplicial complexes where simplices are defined by non-intersecting arcs, called \emph{arc complexes} and have been studied extensively in the literature (see for example \cite{harer85, hatcher91}). We will study an analogous version of this which we call the ``fern complex". \begin{definition} A \emph{fern} is an unordered $\xi$-tuple of arcs that do not intersect except at $*$ and such that their other endpoints are all in the same $\Delta_i$. \end{definition} We will be considering ferns up to isotopy relative to the marked points and $*$. Two or more isotopy classes of ferns are \emph{disjoint} if there exists representative ferns that are disjoint. \begin{definition} \label{def: fern} The \emph{fern complex}, $A_n^\xi= A_n^\xi(S, \Delta_1, \ldots \Delta_n)$, is the simplicial complex such that: \begin{itemize} \item vertices are isotopy classes of ferns; and \item $p$-simplices are collections of $(p+1)$ vertices that are disjoint except at $*$. \end{itemize} \end{definition} See \cref{pic: fern simplex} for examples of simplices in the fern complex of a disk. \begin{center} \begin{figure} \includegraphics[scale=0.26]{fern_vertex1.png} \includegraphics[scale=0.26]{fern_vertex2.png} \caption{A picture of representatives of a vertex (left) and a 2-simplex (right) in $A_4^2(D^2)$. Note that a single fern must have all endpoints of arcs that are not at $*$ in the same subset.} \label{pic: fern simplex} \end{figure} \end{center} \vspace{-10pt} Recall that a space $X$ is $n$-connected if its homotopy groups, $\pi_k(X)$, are trivial for $k \leq n$. The goal of this section is to prove the following connectivity theorem for fern complexes. \begin{theorem} \label{thm: fern connectivity} $A_n^\xi$ is $n-2$ connected. \end{theorem} Our method of proof will be similar to the connectivity proofs found in Section 4 of \cite{wahl12}. In order to prove \cref{thm: fern connectivity}, we will need to study the following related simplicial complex, where we allow ferns to also agree at marked points. \begin{definition} \label{def: fan fern} $FA_n^\xi = FA_n^\xi(S, \Delta_1, \ldots \Delta_n)$ is the simplicial complex such that: \begin{itemize} \item vertices are isotopy classes of ferns; and \item $p$-simplices are collections of $(p+1)$ vertices that are disjoint except possibly at their endpoints. \end{itemize} \end{definition} The following is an analogue to a special case of the main theorem of Hatcher in \cite{hatcher91}, which proves contractibility for certain arc complexes of surfaces. \begin{lemma} \label{lem: fern contractible} $FA_n^\xi$ is contractible. \end{lemma} \begin{proof} Fix a vertex $v$ of $FA_n^\xi$, and a representative fern, which we also call $v$, with arcs going from $*$ to $\Delta_1$ say. Moreover fix an ordering of the arcs of $v = (v_1, \ldots, v_\xi)$. We will show that $FA_n^\xi$ deformation retracts onto $\mathrm{Star}(v)$. Order the interior points of $v$ so that \begin{enumerate} \item $x \prec y$ if $x \in v_i, y \in v_j,$ and $i < j$; \item If $i = j$ then $x\prec y$ if $x$ is closer to $*$ along $v_i$ than $y$. \end{enumerate} Let $\sigma = \langle a_0, \ldots, a_k \rangle$ be a simplex of $FA_n^\xi$. In particular the $a_i$ are $\xi$-tuples of arcs. Choose representatives ferns of $\sigma$ so that $a_0 \cup \ldots \cup a_k$ intersect $v$ minimally. Consider the ferns of $\sigma$ that intersect $v$. Suppose that there are $k$ points of intersection and denote by $g_1, \ldots , g_k$ the germs of arcs that are constituents of ferns corresponding to those intersection points. By germ, we simply mean a small arc segment that intersects $v$. Moreover assume that $g_i \prec g_j$ for $i<j$, where $g_i \prec g_j$ if their corresponding intersection points with $v$ satisfy the inequality $\prec$. The $g_i$ are germs of the arcs of ferns $a_{j_i}$, where it is possible that $j_i = j_{i'}$ for $i \neq i'$ if the arc intersects $v$ more than once. We will now describe a sequence of $k$, $(p+1)$-simplices $r_1(\sigma), \ldots ,r_k(\sigma)$ associated to $\sigma$. Our retraction will then simply be a map that carries $\sigma$ through these simplices. If $\alpha_i$ is an arc intersecting the arc $v_i$ of a fern $v$ at a point $x$, and $x$ is the first intersection point of $\alpha_i$ according to the ordering $\prec$ of $v$, we can define $L(\alpha_i)$ and $R(\alpha_i)$ to be the new path obtained by cutting $\alpha_i$ at $x$ and joining the new endpoints to $*$ by travelling along the left and right hand side of $v_i$. Note that one of the paths of $L(\alpha_i)$ or $R(\alpha_i)$ will not be an arc since it will have one path with both endpoints at $*$. Call $C(\alpha_i)$ the one that is an arc. Now if $a = (\alpha_1, \ldots, \alpha_\xi )$ is a fern, we define $C(a)$ to be the new fern obtained by doing $C$ to the arc intersecting $v$ ``first" according to the ordering $\prec$ on $v$. For example, if $\alpha_1$ had an intersection with $v$ that occurred first according to $\prec$, then $C(a) = (C(\alpha_1), \alpha_2, \ldots , \alpha_\xi)$. Note that $C(\alpha_1)$ will still be a fern whose endpoints still go to the same $\Delta_i$ as $\alpha_1$ since endpoints do not change when doing $C$. We will now use the operator $C$ on ferns to define our sequence of simplices. Let $r_i(\sigma)$ be the $(p+1)$-simplex given by \[ r_i(\sigma) = \langle b_0, \ldots , b_{p+1} \rangle, \] where $b_l = C^{\epsilon_i(l)}(a_l)$ for $l \leq p$, and $b_{p+1} = L^{\epsilon_i(j_i) + 1}(a_{j_i}) = L^{\epsilon_{i+1}(j_i)}(a_{j_i})$, and $\epsilon_i(l)$ is the number $j < i$ such that $g_j$ is a germ of $a_l$. \cref{pic: fern contract} gives an example for $\xi=2$ of this retraction process. \begin{center} \begin{figure} \includegraphics[scale=0.42]{fern_contract.png} \caption[A retraction of a simplex in $FA_{n,\xi}$]{An example (for $\xi = 2$) of germs of ferns of a that intersect $v$, on the left followed by the simplices $r_1, r_2, r_3$ of the retraction in the middle and finally the image of germs on the right. The dotted lines represent the discarded part of arcs during the cutting process.} \label{pic: fern contract} \end{figure} \end{center} \vspace{-30pt} Using barycentric coordinates, a point on $\sigma$ can be identified with the $p$-tuple $(t_0, \ldots, t_p)$, such that $\sum t_i = 1$. We interpret these coordinates as the fern $a_i$ having the weight $t_i$. Assign to the $i$th germ $g_i$ the weight $w_i = t_{j_i}/2$. For $\sum_{j=1}^{i-1} w_j \leq s \leq \sum_{j=1}^{i}w_j$, define $f: I \times FA_n \rightarrow FA_n$ by \[ f(s, [ \sigma, (t_0, \ldots, t_p)]) = [r_i(\sigma), (v_0, \ldots, v_{p+1})],\] where the weight $v_i = t_i$ except for the pair \[ (v_{j_i} , v_{p+1}) = (t_{j_i} - 2(s - \sum_{j=1}^{i-1} w_j), 2(s - \sum_{j=1}^{i-1} w_j)). \] The weight of $(b_{j_i},b_{p+1})$ changes from $(t_{j_i}, 0)$ to $(0, t_{j_i})$ as $s$ goes from $\sum_{j=1}^{i-1} w_j$ to $\sum_{j=1}^i w_j$. This means that the map pushes a face of a simplex through the simplex and onto another face. For $\sum_{i=1}^k w_i \leq s \leq 1$, define $f(s, [\sigma, (t_o, \ldots ,t_p)])$ to be constant, equal to \[f\left(\sum_{i=1}^k w_i, [\sigma, (t_0, \ldots, t_p)]\right).\] In particular $f(1, [\sigma, (t_0, \ldots, t_p)])$ lies in the face of $r_k(\sigma)$ which is in $Star(v)$. The map is continuous since going to a face of $\sigma$ corresponds to a $t_i$ and any corresponding $w_j$ going to zero. \end{proof} We now prove \cref{thm: fern connectivity} by using the contractibility of $FA_n^\xi$. For technical reasons we will need to make use of the following. Given a $p$-simplex $\sigma$ of $FA_n^\xi$ (or $A_n^\xi$), denote by $S_\sigma$ the subspace\footnote{$S_\sigma$ may not be a surface (with boundary) because of the point $*$. On the other hand, this will not be a problem since many of our constructions will still work in this setting.} $S - (\sigma - *) \subset S$. That is, it is the surface, $S$, with a representative fern of $\sigma$ removed, except for the point $*$ on the boundary. For spaces of the form $S_\sigma$, we define $A_n^\xi(S_{\sigma}, \Delta_1, \ldots, \Delta_n)$ and $FA_n^\xi(S_\sigma, \Delta_1, \ldots, \Delta_n)$ as in \cref{def: fern} and \cref{def: fan fern} , with $S$ replaced with $S_\sigma$ in both definitions, where the marked points of $S_\sigma$ are inherited from $S$ and the point $* \in S_\sigma$ is the point $* \in \partial S$. The same arguments as in the proof of \cref{lem: fern contractible} can also be used to show that $FA_n^\xi(S_\sigma, \Delta_1, \ldots, \Delta_n)$ is contractible, which we will make use of in the following proof of \cref{thm: fern connectivity}. \begin{proof}[Proof of \cref{thm: fern connectivity}] Recall we are trying to prove that $A_n^\xi(S, \Delta_1, \ldots, \Delta_n)$ is $n-2$ connected. We will actually prove the theorem for $S$ a surface or of the form $S_\sigma$, where $\sigma \in (A_n^\xi)_p$ is a $p$-simplex. The proof will proceed by induction on $n$. The base case of our induction requires us to show that $A_1^\xi(S, \Delta_1)$ is nonempty which is true as long as $S$ is connected. For the inductive step, let $k \leq n -2$. Consider a map \[ f: S^k \rightarrow A_n^\xi. \] We want to show that $f$ factors through a $(k+1)$-disk. By contractibility of $FA_n^\xi$ we have a commutative diagram \[ \xymatrix{ S^k \ar[r]^{f\;\;} \ar@{^{(}->}[d] & A_n^\xi \ar@{^{(}->}[d] \\ D^{k+1} \ar[r]^{\hat{f}\;\;} & FA_n^\xi \\ } \] By simplicial approximation, we can triangulate $S^k$ and $D^{k+1}$ and take our maps to be simplicial. We want to deform $\hat{f}$ so that its image lies in $A_n^\xi$. A fern is \emph{coloured by $\Delta_i$} if the endpoints of its arcs are at $\Delta_i$. Call a simplex $\sigma = \langle a_0, \ldots, a_p \rangle \in D^{k+1}$ bad if the colour of each $\hat{f}(a_i)$ already appears as a colour of one of the $\hat{f}(a_j)$ for $i \neq j$. Let $\sigma \in D^{k+1}$ be a bad simplex of maximal dimension, say $p$. Then $\hat{f}$ restricts to a map \[ \hat{f} |_{\mathrm{Link}(\sigma)} : \mathrm{Link}(\sigma) \rightarrow J_\sigma := A_{n'}^\xi(S_\sigma, \Delta'_0, \ldots, \Delta'_{n'}) \] where the $\Delta'_i$ are the remaining partitions of marked points not at the endpoints of ferns in $\hat{f}(\sigma)$. The image of the map $\hat{f}$ restricts to $A_{n'}^\xi(S_\sigma, \Delta'_0, \ldots, \Delta'_{n'})$ because if $\tau \in \mathrm{Link}(\sigma)$ was bad, then $\sigma * \tau$ would be a bad simplex in $D^{k+1}$ of larger dimension than $\sigma$ contradicting maximality of $\sigma$. Now $n' \geq n - \left\lfloor (p+1)/2 \right\rfloor$, since a bad $p$-simplex can use up at most $\left\lfloor (p+1)/2 \right\rfloor$ $\Delta_i$'s. Also note that $p \geq 1$ since there are no bad vertices. By induction the connectivity of $J_\sigma$ is \begin{align*} \mathrm{conn}(J_\sigma) &= n' - 2\\ &\geq n - \left\lfloor \frac{p+1}{2} \right\rfloor - 2 \\ &\geq n - 2 - p + \left\lfloor \frac{p-1}{2} \right\rfloor\\ & \geq k - p. \end{align*} Since the link of $\sigma$ is a $k-p$ sphere, we have a commutative diagram \[ \xymatrix{ \mathrm{Link}(\sigma) \ar[r] \ar@{^{(}->}[d] & J_\sigma \ar[r] & A^\xi_n(S, \Delta_0, \ldots \Delta_n) \\ K \ar[ur]_{\hat{f}'} & & } \] where $K$ is a $(k-p+1)$-disk with boundary $\partial K = \mathrm{Link}(\sigma)$ and the right map is the map that identifies arcs on $S'$ with arcs on $S$. In the triangulation of $D^{k+1}$, replace the $(k+1)$-disk, $\mathrm{Star}(\sigma) = \sigma * \mathrm{Link}(\sigma)$ with $\partial \sigma * K$. We can do this since $\partial \sigma *K$ and $\mathrm{Star}(\sigma)$ have the same boundary. Modify the map $\hat{f}$ by \[ \hat{f} * \hat{f}': \partial \sigma * K \rightarrow FA_n^\xi(S, \Delta_0, \ldots, \Delta_n). \] New simplices in $\partial \sigma * K$ are of the form $\tau = \alpha * \beta$, where $\alpha$ is a proper face of $\sigma$ and $\beta$ is mapped to $J_\sigma$. Thus, if $\tau$ is a bad simplex in $\partial \sigma * K$ then $\tau = \alpha$ since ferns of $\hat{f}'(\beta)$ do not share colours with other ferns of $\hat{f}'(\beta)$ or $\hat{f}(\alpha)$, so cannot contribute to a bad simplex. Since $\alpha$ is a proper face of $\sigma$, we have decreased the number of top dimensional bad simplices. The result follows by inducting on top dimensional bad simplices.\end{proof} \section{The stabilisation map} \label{sec: stab} In this section, we will define the stabilisation maps of \cref{thm: surface braid stability}. The basic example that one should have in mind is the stabilisation map for braid groups $\mathrm{stab}: \mathrm{Br}_n \rightarrow \mathrm{Br}_{n+1}$ which adds an unbraided strand. Let $S$ be the interior of a surface $\bar{S}$ with boundary $\partial \bar{S}$. Pick a point $b \in \partial_0 \subset \partial \bar{S}$, where $\partial_0$ is a boundary component of $\partial \bar{S}$. On the level of ordered $\xi$-configuration spaces, we define a map \[ s :\mathrm{PConf}_n^\xi(S) \rightarrow \mathrm{PConf}_{n+1}^\xi(S'), \] where $S'$ is obtained from $\bar{S}$ by adding a collar neighbourhood, $\partial_0 \times I,$ around $\partial_0$. Let $\bm{q}$ be the subset of the collared neighbourhood given by \[ \bm{q} = \{ (b, \frac{1}{\xi+1}), \ldots, (b, \frac{\xi}{\xi+1}) \}.\] Then $s$ is defined by inserting this new subset $(\bm{p}_1, \ldots, \bm{p}_m) \mapsto (\bm{p}_1, \ldots, \bm{p}_m, \bm{q}).$ Picking an isomorphism $f: S \cong S'$, with support in a small neighbourhood of $\partial_0$, we get a map $f \circ s :\mathrm{PConf}_n^\xi(S) \rightarrow \mathrm{PConf}_{n+1}^\xi(S').$ Define the \emph{stabilisation map} for $\xi$-configuration spaces to be the map \[ \mathrm{stab}: \mathrm{Conf}_n^\xi(S) \rightarrow \mathrm{Conf}_{n+1}^\xi(S)\] obtained by quotienting $f \circ s$ by the symmetric group action, that is, it is the map \[ \mathrm{stab} : \frac{\mathrm{PConf}_n^\xi(S)}{\Sigma_n} \rightarrow \frac{\mathrm{PConf}_{n+1}^\xi(S)}{\Sigma_{n+1}}.\] \begin{remark} The stabilisation map for $\xi$-configurations spaces that we have defined corresponds to the map on the level of surface braid groups $\mathrm{stab}: \mathrm{Br}_n^\xi(S) \rightarrow \mathrm{Br}_{n+1}^\xi(S)$ which adds $\xi$ trivial strands. In particular, $\mathrm{Br}_n^\xi(S)$ includes into $\mathrm{Br}_{n+1}^\xi(S)$ as the inclusion of a stabiliser of a fern ending at the added marked points of $\mathrm{Conf}_{n+1}^\xi(S)$. \end{remark} \begin{figure} \begin{center} \includegraphics[scale=0.40]{xi_stab.png} \caption[The stabilisation map for $\xi$-configurations]{The stabilisation map $\mathrm{stab}: \mathrm{Conf}_n^\xi(S) \rightarrow \mathrm{Conf}_{n+1}^\xi(S)$ for $\xi=2$ and $n=2$}\label{fig: stab} \end{center} \end{figure} \section{Homological stability for open surfaces} \label{sec: open} The goal of this section is to prove \cref{thm: surface braid stability}. For convenience we recall the statement of the theorem. \begin{theorem}\label{thm: surface braid stability} Let $S$ be a surface that is the interior of a surface with boundary. The map \[ \mathrm{stab}_* : H_*(\mathrm{Br}_n^\xi(S); \mathbb{Z}) \rightarrow H_*(\mathrm{Br}_{n+1}^\xi(S); \mathbb{Z}) \] is an isomorphism for $2* \leq n$. \end{theorem} \begin{proof} For brevity, we write $\mathrm{Br}_n^\xi$ for $\mathrm{Br}_n^\xi(S)$. By Theorem 5.1 of \cite{hw10}, it suffices to check the following conditions to prove that we have isomorphisms in the range $* \leq n/2-1$ and surjections for $* \leq n/2$. \begin{enumerate} \item The action of $\mathrm{Br}_n^\xi$ on $A_n^\xi$ is transitive on vertices; the stabiliser of each simplex fixes the simplex point wise; and $H_i(A_n^\xi / \mathrm{Br}_n^\xi) = 0$ for $1 \leq i \leq n - 2$. \item The subgroup of $\mathrm{Br}_n^\xi$ fixing a $p$-simplex pointwise is conjugate to $\mathrm{Br}_{n - \xi(k + 1)}^\xi$ for some $k \leq p$. \item For each edge of $A_n^\xi$ with vertices $v$ and $w$ there exists an element of $\mathrm{Br}_n^\xi$ that takes $v$ to $w$ and that commutes with elements of $\mathrm{Br}_n^\xi$ that leave the edge fixed pointwise. \end{enumerate} For the first condition, note that a vertex of $A_n^\xi$ can be identified uniquely with $\xi$-simplices in $A_{n\xi}$, where $A_{n\xi} := A_{n\xi}^1$ is the arc complex of a disk. It is well known that the action of $\mathrm{Br}_n$ on $(A_{n\xi})$ is transitive. Moreover, if the simplices have all arcs going to the same $\Delta_i$, it is possible to choose the element of $\mathrm{Br}_n$ take one simplex to another to preserve the $\Delta_i$ so that it is in $\mathrm{Br}_n^\xi$, which shows that the action of $\mathrm{Br}_n^\xi$ on $A_n^\xi$ is transitive on vertices. The stabiliser of a simplex fixes the simplex pointwise as the order in which arcs appear at the base point cannot be changed. To see that $H_i(A_n^\xi / \mathrm{Br}^\xi_n)= 0$ for $1 \leq i \leq n-2$ we use an argument from the proof of Lemma 3.3 in \cite{harer85}. The aim is to consider the chain complex $\mathbb{Z}(A_n^\xi / \mathrm{Br}^\xi_n)_p$ (we will describe the differentials momentarily) and construct a chain nullhomotopy. It is possible to identify the orbit of a $p$-simplex $\sigma$ so that there is a correspondence between $A_n^\xi / \mathrm{Br}^\xi_n$ and orderings of the multiset $\{ 0, \ldots, 0, \ldots, p, \ldots, p\}$ where the first $0$ appears before the first $1$, the first $1$ appears before the first $2$ and so on and each number appears $\xi$ times. For example $(0,1,1,2,0,2)$ is an allowed ordering but $(0,2,0,1,1,2)$ is not. To see this correspondence, we put an ordering on the vertices of $A_n^\xi$ so that for two ferns $\sigma$ and $\tau$ in minimal position, $\sigma < \tau$ if the leftmost arc of $\sigma$ is further to the left than the leftmost arc of $\tau$. Here by leftmost we refer to the left to right ordering of the tangent vectors of the arcs at the base point $*$. With this ordering, $A_n^\xi$ is an ordered simplicial complex. Let $\langle a_0, \ldots a_p \rangle$ be a $p$-simplex in $A_n^\xi$. Labelling all the arcs of $a_i$ by the number $i$ and reading these labels off from left to right at $*$ gives rise to the corresponding ordering of the multiset $\{ 0, \ldots, 0, \ldots, p, \ldots, p\}$. Two ferns that give rise to different orderings of multisets are in different $\mathrm{Br}_n^\xi$ orbits since the action of $\mathrm{Br}_n^\xi$ cannot change the ordering of arcs at $*$. On the other hand, simplices giving rise to the same ordering are in the same orbit by an argument similar to the transitivity argument for vertices. See \cref{pic: fern hom} for a picture of two ferns in the same orbit. \begin{center} \begin{figure} \includegraphics[scale=0.28]{fern_orbit_2.png} \includegraphics[scale=0.28]{fern_orbit_3.png} \caption{Two ferns in the orbit labelled by $(0, 1, 1, 2, 0 , 2)$} \label{pic: fern hom} \end{figure} \end{center} \vspace{-25pt} For $ 0 \leq i \leq p$, with the above description of simplices in the ordered fern complex $A_n^\xi$, the face maps of $A_n^\xi$ as an ordered simplicial complex are given by \[ f_i (\sigma) = \mathrm{forget}_i(\sigma) \] where $\mathrm{forget}_i$ forgets the $i$'s in $\sigma$ and then subtracts $1$ from all the numbers greater than $i$. For example, $f_1(0, 0, 2, 3, 2, 1,1, 3) = (0,0, 1, 2, 1, 2)$. Thus the differentials of the chains on $A_n^\xi$, $d : \mathbb{Z}(A_n^\xi / \mathrm{Br}_n^\xi)_p \rightarrow \mathbb{Z}(A_n^\xi / \mathrm{Br}_n^\xi)_{p-1}$ are given by the map \[ d(\sigma) = \sum_j (-1)^j f_j(\sigma) . \] For $0 \leq p \leq n - \xi$, define $D : (A_n^\xi / \mathrm{Br}^\xi_n)_p \rightarrow (A_n^\xi / \mathrm{Br}_n^\xi)_{p+1}$ by taking a $p$-simplex $\sigma$, adding one to every entry and then putting $\xi$ zeroes in front. For example $D(0 , 0 ,2 , 2, 1, 1) = (0, 0 , 1, 1, 3, 3, 2, 2)$. We now have that $Dd + d D = id$ so the identity map is chain homotopic to zero in the range $0 \leq p \leq n - \xi$. Thus $H_i(A_n^\xi / \mathrm{Br}^\xi_n) = 0 $ for $1 \leq i \leq n-\xi$. For the second condition, we get the conjugation by identifying the stabiliser of a $p$-simplex with the $\xi$-braid group that acts on the cut surface where we cut along the arcs of the $p$-simplex. Finally for the last condition, the action of $\mathrm{Br}^\xi_n$ that takes a vertex $v$ to a vertex $w$ is supported on a tubular neighbourhood of $v \cup w$. Theorem 5.1 of \cite{hw10} now implies that we have isomorphisms in the range $* \leq n/2 - 1$ and an epimorphism for $* \leq n/2$. In \cref{prop: surface braid transfer}, we will in fact show that the maps \[ \mathrm{stab}: \mathrm{Br}_n^\xi(S) \rightarrow \mathrm{Br}_{n+1}^\xi(S) \] are always injective in homology and this gives the full result.\end{proof} \section{Homological stability for closed surfaces} \label{sec: braid stability closed} When $S$ is a closed surface, it is not possible to define the stabilisation maps that we have been using to prove our homological stability theorems. The main issue is that we no longer have a boundary from which to push in new points. Indeed, even for the usual configuration space, integral homological stability for configuration spaces of closed manifolds does not hold (see for example \cite{fvb62} from which one can compute $H_1(\mathrm{Conf}_n(S^2); \mathbb{Z}) \cong \mathbb{Z}/(2n-2)$). On the other hand if one considers rational homology, then Church and Randal-Williams \cite{church12, orw13} prove that homological stability for configuration spaces of closed manifolds holds rationally. In this section, we will prove the analogous result for closed surfaces and $\xi$-configuration spaces (or equivalently $\xi$-braid groups). We use a technique of Randal-Williams in \cite{orw13} where he proves homological stability for configuration spaces of closed manifolds, using homological stability for open ones. This will involve defining transfer maps which realise the homology isomorphisms. These maps will be more naturally defined in terms of $\xi$-confiugration spaces rather than $\xi$-braids. We define these transfer maps as follows. For $m < n$, let $\mathrm{Conf}_{n,m}^\xi(S)$ be the $\xi$-configuration space where the $n$ subsets have been partitioned into subsets of size $m$ and $n-m$. There is a covering map \[ p : \mathrm{Conf}_{n,m}^\xi(S) \rightarrow \mathrm{Conf}_n^\xi(S) \] obtained by forgetting the partitioning. There is a transfer map \[ t_{n,m} :H_*(\mathrm{Conf}_n^\xi(S) ;\mathbb{Z}) \rightarrow H_*(\mathrm{Conf}_{n,m}^\xi(S); \mathbb{Z}) \rightarrow H_*(\mathrm{Conf}_m^\xi(S); \mathbb{Z}) \] where the first map is the transfer map associated to the covering $p$, and the second is the map induced by \[ f: \mathrm{Conf}_{n,m}^\xi(S) \rightarrow \mathrm{Conf}_m^\xi(S),\] which forgets down to $m$ subsets. We use the notation $t_n$ to denote $t_{n, n-1}$ \begin{remark} There is also a way to describe the transfer map on the level of spaces using the identification $H_*(X) \cong \pi_*(\mathrm{Sym}^\infty(X))$ from the Dold-Thom theorem. Given an $n$-fold covering $\pi: E \to B$, there is a map $\tau : B \rightarrow \mathrm{Sym}^n(E)$, given by sending a point in $b$ to $\pi^{-1}(b)$. The transfer map on homology is defined as the map $H_*(B) \xrightarrow{\tau_*} H_*(E)$, given by the induced map on $\pi_*$ of $\mathrm{Sym}^\infty (\tau) : \mathrm{Sym}^\infty B \rightarrow \mathrm{Sym}^\infty (E)$, where we have identified $\mathrm{Sym}^\infty(\mathrm{Sym}^n(E))$ with $\mathrm{Sym}^\infty(E)$ and then used the Dold-Thom theorem. We will make use of this description of the transfer map in the proof of \cref{thm: surface braid closed}. \end{remark} \begin{proposition}\label{prop: surface braid transfer} Let $S$ be a surface that is the interior of a surface with boundary. The map \[ (\mathrm{stab}_n)_* : H_*(\mathrm{Conf}_n^\xi(S) ; \mathbb{Z}) \rightarrow H_*(\mathrm{Conf}_{n+1}^\xi(S); \mathbb{Z}) \] is always split injective. Moreover, the transfer map $t_n$ is a rational isomorphism whenever $\mathrm{stab}_{n-1}$ is. \end{proposition} \begin{proof} For brevity let $s_n$ denote the stabilisation map $(\mathrm{stab}_n)_*$. From the definition of the transfer, we see that the maps $s$ and $t$ satisfy the relations \[ t_j \circ s_{j-1} = s_{j-2} \circ t_{j-1} + id. \] More generally, they satisfy \[ t_{j,k}\circ s_{j-1} = s_{j-2} \circ t_{j-1, k-1} + t_{j-1, k}. \] Furthermore, \[ t_{k+1} \circ \cdots \circ t_j = (j-k)!t_{k,j}. \] Letting $A_j = H_i(\mathrm{Conf}_n^\xi(S); \mathbb{Z})$ and $B_j := \mathrm{coker}(s_{j-1}),$ we are now in the situation of \cite[Lemma 2]{Dold62}, which implies that the $s_j$ are split injective and that \[ t_{j+1} \circ s_j \] is multiplication by a nonzero constant. On homology, this is rationally an isomorphism so $t_{j+1}$ is an isomorphism whenever $s_j$ is an isomorphism which gives the desired bound. \end{proof} We have now completed the proof of \cref{thm: surface braid stability}, using the transfer maps to show that $\mathrm{stab}: \mathrm{Conf}_n^\xi(S) \rightarrow \mathrm{Conf}_{n+1}^\xi(S)$ is always injective and is rationally inverse to the transfer map in a range. Since the transfer map does not require us to add points to our surface, it makes sense to talk about the transfer map even for closed manifolds. In this way, we get a map $t_n:H_*(\mathrm{Conf}_n^\xi(S)) \rightarrow H_*(\mathrm{Conf}_{n-1}^\xi(S))$ and we can ask when this map is an isomorphism. In general, stability does not hold with integral coefficients. However, if we instead work with rational coefficients then we can prove the following. \begin{theorem} \label{thm: surface braid closed} Let $S$ be a surface, open or closed. \[ t : H_k(\mathrm{Conf}_n^\xi(S) ; \mathbb{Q}) \rightarrow H_k(\mathrm{Conf}_{n-1}^\xi(S) ; \mathbb{Q}) \] is an isomorphism for $2k \leq n-1 $. \end{theorem} The proof of \cref{thm: surface braid closed} will be similar to the proof of Proposition 9.4 in \cite{orw13}. We will make use of the following construction. Let \[ D_n^\xi(S)_i := \{( \underline{c}, p_0, \ldots, p_i) \in \mathrm{Conf}^\xi_n(S) \times S^{i+1} \; \vert \; p_j \not\in \underline{c } \mbox{ and } p_j \neq p_k \}. \] The notation $p_j \not\in \underline{c}$ means the points $p_j$ do not lie in any of the subsets that make up $\underline{c}$. There are face maps $\partial_j : D_n^\xi(S)_i \rightarrow D_n^\xi(S)_{i-1}$ obtained from forgetting $p_j$. Moreover there is an augmentation map $\epsilon: D_n^\xi(S)_\bullet \rightarrow \mathrm{Conf}_n^\xi(S)$ which forgets all the $p_j$ so that $D_n^\xi(S)_\bullet$ is an augmented semi-simplicial space. We can similarly define the augmented semi-simplicial space \[ D_{n,1}^\xi(S)_i := \{( \underline{c}, p_0, \ldots, p_i) \in \mathrm{Conf}^\xi_{n,1}(S) \times S^{i+1} \; \vert \; p_j \not\in \underline{c } \mbox{ and } p_j \neq p_k \}. \] We will make use of the following lemma which follows from the proof of Lemma 2.1 in \cite{orw10}. \begin{lemma} Let $\norm{X_\bullet} \rightarrow X$ be an augmented (semi-)simplicial space, $f: X_n \rightarrow X$ be the unique face map and let $x \in X$. If $f$ is a Serre fibration, then \[ \norm{ f^{-1}(x)_\bullet} \rightarrow \norm{X_\bullet} \rightarrow X \] is a homotopy fibre sequence. \label{lem: fibre sequence} \end{lemma} \begin{lemma} \label{lem: braid weak equiv}The maps $\norm{D_n^\xi(S)_\bullet} \to \mathrm{Conf}_n^\xi(S)$ and $\norm{D_{n,1}^\xi(S)_\bullet} \to \mathrm{Conf}_{n,1}^\xi(S)$ are weak equivalences. \end{lemma} \begin{proof} We prove that the first map is a weak equivalence. By \cref{lem: fibre sequence} the map $ \norm{D_n^\xi(S)_\bullet} \rightarrow \mathrm{Conf}_n^\xi(S) $ is a homotopy fibre sequence with fibre over a $\xi$-configuration $\underline{c}$ given by $\norm{F(S - \underline{c})_\bullet}$, where $F(S - \underline{c})_\bullet$ is the semi-simplicial space whose $i${th} term consists of unordered $(i+1)$-tuples of distinct points in $S - \underline{c}$. We will show that $\norm{F(S - \underline{c})_\bullet}$ is contractible. By taking small neighbourhoods of the points in $\underline c$, we can find a closed surface $S' \subset (S - \underline{c})$ which is homotopy equivalent to $(S - \underline{c})$ with some point $x \in (S - \underline{c}) - S'$. Now suppose we have map $f : S^k \to \norm{F(S - \underline{c})_\bullet}$. By the previous homotopy equivalence, we can deform $f$ so that $x$ does not lie in its image. Moreover, by simplicial approximation, we can find a PL triangulation of $S^k$ and take $f$ to be simplicial. Now, we can fill in $f$ by defining a map $\hat{f} : \mathrm{cone}(S^k) \to \norm{F(S - \underline{c})_\bullet}$ that sends the cone point to $x$. Therefore $\norm{F( S - \underline{c})_\bullet}$ is contractible and so the map $\norm{D_n^\xi(S)_\bullet} \rightarrow \mathrm{Conf}_n^\xi(S)$ is an equivalence. The argument that $\norm{D_{n,1}^\xi(S)_\bullet} \to \mathrm{Conf}_{n,1}^\xi(S)$ is a weak equivalence is similar. \end{proof} There are semi-simplicial maps \[ D_n^\xi(S)_\bullet \leftarrow D_{n,1}^\xi(S)_\bullet \rightarrow D_{n+1}^\xi(S)_\bullet \] modelled on the maps for $\xi$-configurations, $\mathrm{Conf}_n^{\xi}(S) \leftarrow \mathrm{Conf}_{n,1}^\xi(S) \rightarrow \mathrm{Conf}_{n+1}^\xi(S)$. \begin{lemma} \label{lem: braid resolution} There is a map of semi-simplicial spaces \[ D_n^\xi(S)_\bullet \xrightarrow{\tau} \mathrm{Sym}^n_{fib}(D_{n,1}^\xi(S)_\bullet) \rightarrow \mathrm{Sym}^n_{fib} (D^\xi_{n+1}(S)_\bullet) \] which induces a map \[ \norm{D_n^\xi(M)_\bullet} \xrightarrow{\tau} \mathrm{Sym}^n(\norm{D_{n,1}^\xi(S)_\bullet}) \rightarrow \mathrm{Sym}^n(\norm{D_{n+1}^\xi(S)_\bullet}) \] on geometric realisations. The terms in these maps are described in the proof. \end{lemma} \begin{proof} There are fibration sequences given by \[ \mathrm{Conf}_n^\xi(S - \mbox{ $i+1$ points }) \rightarrow D_n^\xi(S)_i \xrightarrow{\pi} \mathrm{PConf}_{i+1}(S) \] where $\pi$ is the map that sends $(\underline{c}, p_0, \ldots, p_i) \mapsto (p_0, \ldots, p_i)$. Let $\mathrm{Sym}^n_{fib}(D_n^\xi(S)_i)$ denote the $n$-fold fibrewise symmetric product with respect to this fibration. The maps $D_{n,1}^\xi(S)_i \rightarrow D_{n+1}^\xi(S)$ are fibrewise $n$-fold coverings over $\mathrm{PConf}_{i+1}(S)$ which give rise to transfer maps \[ \tau : D_{n+1}^\xi(S)_i \rightarrow \mathrm{Sym}^n_{fib} (D_{n,1}^\xi(S)_i).\] One can check that $\tau$ commutes with the face maps and so defines a transfer $\tau$ on geometric realisations as in the statement of the lemma. \end{proof} We can now prove \cref{thm: surface braid closed} \begin{proof}[Proof of \cref{thm: surface braid closed}] Associated to a semi-simplicial space is a spectral sequence which computes the homology of its geometric realisation in terms the the homology of its levels. Therefore \cref{lem: braid resolution} gives rise to a map of spectral sequences converging to the transfer map \[ t: H_*( \mathrm{Conf}_n^\xi(S) ; \mathbb{Q}) \rightarrow H_*(\mathrm{Conf}_{n-1}^\xi(S); \mathbb{Q}) \] where we have used \cref{lem: braid weak equiv} to identify $\mathrm{Conf}_n^\xi(S) \simeq \norm{D_n^\xi(S)}$ and the Dold-Thom theorem to get a map from $H_*(\norm{D_n^\xi(S)_\bullet} ; \mathbb{Q}) \rightarrow H_*(\norm{D_{n-1}^\xi(S)_\bullet}; \mathbb{Q})$. The map on $E^1_{pq}$ terms of the spectral sequence is \[ H_q(D_n^\xi(S)_p ; \mathbb{Q}) \rightarrow H_q(D_{n-1}^\xi(S)_p ; \mathbb{Q})\] and is induced by $D_n^\xi(M)_i \rightarrow \mathrm{Sym}^n_{fib}(D_{n,1}(S)_i) \rightarrow \mathrm{Sym}^n_{fib}(D_{n-1}^\xi(S)_p)$. From the fibration in the proof of \cref{lem: braid resolution}, there is a Serre spectral sequence converging to this map. The map on $E^2_{pq}$ terms of this Serre spectral sequence is \begin{align*} H_p(\mathrm{PConf}_i(S); H_q( \mathrm{Conf}_n^\xi(&S - \mbox{$i+1$ points}); \mathbb{Q})) \rightarrow \\ & H_p(\mathrm{PConf}_{i}(S); H_q( \mathrm{Conf}_{n-1}^\xi(S - \mbox{$i+1$ points}); \mathbb{Q})). \end{align*} On coefficients, it is induced by the transfer map \[t: H_q( \mathrm{Conf}_n^\xi(S - \mbox{$i+1$ points}); \mathbb{Q}) \rightarrow H_q( \mathrm{Conf}_{n-1}^\xi(S - \mbox{$i+1$ points}); \mathbb{Q}). \] Since $(S - \mbox{$i + 1$ points})$ is an open surface, \cref{prop: surface braid transfer} implies that this is an isomorphism for $q \leq (n-1)/2$. Thus the map $H_q( D_n^\xi(S)^p; \mathbb{Q}) \rightarrow H_q( D_{n-1}^\xi(S)^p; \mathbb{Q})$ is an isomorphism in this range. In particular, \[ t_n: H_*(\mathrm{Conf}_n^\xi(S) ;\mathbb{Q}) \rightarrow H_*(\mathrm{Conf}_{n-1}^\xi(S) ; \mathbb{Q}) \] is an isomorphism for $* \leq (n-1)/2$. \end{proof} \bibliographystyle{alpha}
\section{Introduction}\label{intro} The launch of {\it Fermi Gamma-Ray Space Telescope} (hereafter {\it Fermi}) in 2008 has led to the discovery of $\gamma$-ray emission in a large number of blazars, comparable to that found by the {\it Energetic Gamma-Ray Experiment Telescope} \citep[{\it EGRET};][]{1993ApJS...86..629T}, owing to its large sensitivity and wide energy coverage. This has clearly demonstrated that blazars dominate the population of $\gamma$-ray emitters for which classification is known \citep{2012ApJS..199...31N}. Interestingly, {\it Fermi} has also observed variable $\gamma$-ray emission from some radio-loud narrow line Seyfert 1 (RL-NLSy1) galaxies \citep{2009ApJ...699..976A,2009ApJ...707L.142A,2012MNRAS.426..317D,2011MNRAS.413.2365C}. First identified as a new class of active galactic nuclei (AGN) by \citet{1985ApJ...297..166O}, NLSy1 galaxies have narrow Balmer lines (FWHM (H$_{\beta}$) $<$ 2000 km s$^{-1}$), weak [O~{\sc iii}] ([O~{\sc iii}]/H$_{\beta} <$ 3) and strong optical Fe~{\sc ii} lines \citep{1985ApJ...297..166O,1989ApJ...342..224G}. They also posses soft X-ray excess \citep{1996A&A...305...53B,1996A&A...309...81W,1999ApJS..125..297L}, relatively low black hole (BH) mass ($\sim$ 10$^{6}-10^{8}$ \ensuremath{M_{\odot}}), high accretion rate \citep{2000ApJ...542..161P,2000NewAR..44..419H,2004ApJ...606L..41G,2006ApJS..166..128Z,2012AJ....143...83X}, and rapid X-ray flux variations \citep{1995MNRAS.277L...5P,1999ApJS..125..317L}. They also exhibit the radio-quiet/radio-loud dichotomy \citep{2000ApJ...543L.111L} and $\sim$ 7\% NLSy1 galaxies are found to be radio-loud \citep{2006AJ....132..531K}, a lesser fraction when compared to the $\sim$ 15\% found in quasar population \citep{1995PASP..107..803U}. Some of these RL-NLSy1 galaxies show compact core-jet structure, high brightness temperature and superluminal behavior \citep{2006AJ....132..531K,2006PASJ...58..829D}. Kilo-parsec scale radio structures are also reported in some RL-NLSy1 galaxies \citep{2008A&A...490..583A,2012ApJ...760...41D}. All these characteristics indicated the presence of aligned relativistic jets in RL-NLSy1 galaxies and the detection of $\gamma$-ray emission from five RL-NLSy1 galaxies by the Large Area Telescope (LAT) onboard {\it Fermi} provided the confirmation \citep{2009ApJ...699..976A,2009ApJ...707L.142A,2012MNRAS.426..317D}. Recently, high and variable optical polarization and rapid optical/infra-red flux variations are observed in some of these $\gamma$-ray emitting NLSy1 ($\gamma$-NLSy1) galaxies \citep{2011PASJ...63..639I,2010ApJ...715L.113L,2013MNRAS.428.2450P,2012ApJ...759L..31J,2013ApJ...775L..26I}. Modeling the spectral energy distribution (SED) of $\gamma$-NLSy1 galaxies led to the conclusion that they posses the characteristic double hump SED similar to blazars and the physical properties of these sources are intermediate to the other two members of the blazar family, namely BL Lac objects and flat spectrum radio quasars (FSRQs) \citep{2009ApJ...707L.142A,2012MNRAS.426..317D,2011MNRAS.413.1671F,2013ApJ...768...52P}. Also, \citet{2010ASPC..427..243F} have shown that one of these $\gamma$-NLSy1 galaxies, PMN J0948+0022 occupy an intermediate position in the well known blazar sequence \citep{1998MNRAS.299..433F}. Though, many properties of $\gamma$-NLSy1 galaxies are similar to that of blazars, there exist characteristic differences as well. For example, blazars are known to be hosted by massive elliptical galaxies \citep{2009arXiv0909.2576M}, whereas, the $\gamma$-NLSy1 galaxies, similar to other NLSy1 galaxies, are likely to be hosted by spiral galaxies with relatively low-mass BHs at their centers. Therefore, detailed studies of these $\gamma$-NLSy1 galaxies will enable us understand the physical properties of relativistic jets at different mass and accretion rate scales. Also, a clear understanding of the shape of the observed $\gamma$-ray spectrum can help us place constraints on the various models proposed to explain the high energy component present in broadband SED of these sources. In this work, we present $\gamma$-ray flux variability properties and $\gamma$-ray spectral properties of five $\gamma$-NLSy1 galaxies using $\sim$ 5 years of {\it Fermi}-LAT data. Wherever possible, we also study their $\gamma$-ray spectral behavior in different activity states. The motivation here is to see the similarities and/or differences in both the $\gamma$-ray spectra as well as the $\gamma$-ray flux variability of these sources with respect to the blazar class of AGN that emits copiously in the $\gamma$-ray band. The paper has been organized as follows. Sample selection is discussed in Section~\ref{sample}. Section~\ref{data_red} is devoted to data reduction and analysis procedures. We discuss the results in Section~\ref{result} followed by our conclusions in Section~\ref{conclusion}. Throughout the work, we adopt $\Omega_{m}$ = 0.27, $\Omega_{\varLambda}$ = 0.73 and {\it H}$_{0}$ = 71 km s$^{-1}$ Mpc$^{-1}$. \section{Sample}\label{sample} As of now (2014 September 01), five RL-NLSy1 galaxies are found to be emitting in the $\gamma$-ray band by {\it Fermi}-LAT with high significance, having test statistic (TS) $>$ 25 \citep[$\sim$ 5$\sigma$;][]{1996ApJ...461..396M}. General information about these five $\gamma$-NLSy1 galaxies is given in Table~\ref{general}. In Table~\ref{general}, the values of right ascension, declination, redshift, and the apparent V-band magnitude are taken from \citet{2010A&A...518A..10V}. The given radio spectral indices are calculated using the 6 cm and 20 cm flux densities found in \citet{2010A&A...518A..10V} ($S_{\nu} \propto \nu^{\alpha}$) and the values of radio loudness parameter R\footnote{defined as the ratio of flux density in 5 GHz to that in optical {\it B}-band} are taken from \citet{2011nlsg.confE..24F}. \section{Data reduction and Analysis}\label{data_red} The observational data used in this work were collected by {\it Fermi}-LAT \citep{2009ApJ...697.1071A} over the first five years of its operation, from 2008 August 05 (MJD 54,683) to 2013 November 01 (MJD 56,597). For the analysis of the data, standard procedures described in the {\it Fermi}-LAT documentation are used. We use {\tt ScienceTools v9r32p5} along with the post-launch instrument response functions (IRFs) P7REP\_SOURCE\_V15. In the energy range of 0.1$-$300 GeV, only events belonging to the SOURCE class are selected. To avoid contamination from Earth limb $\gamma$-rays, source events coming from zenith angle $<$ 100$^{\circ}$ are extracted. A filter of ``\texttt{DATA$\_$QUAL==1}'', ``\texttt{LAT$\_$CONFIG==1}'', and ``\texttt{ABS(ROCK\_ANGLE)$<$52}'' is used to select the good time intervals. To model the background, galactic diffuse emission component (gll\_iem\_v05.fits) and an isotropic component iso\_source\_v05.txt\footnote{http://fermi.gsfc.nasa.gov/ssc/data/access/lat/BackgroundModels.html} are considered. While fitting the five year average spectrum, the normalization of components in the background model as well as the photon index of the galactic diffusion model are allowed to vary freely. We fix these parameters to the values obtained from the fitting during the subsequent time series analysis. To evaluate the significance of the $\gamma$-ray signal we use the maximum-likelihood (ML) test statistic TS = 2$\Delta$ log(likelihood) between models with and without a point source at the position of the source of interest. We use the binned likelihood method to generate the average $\gamma$-ray spectrum as well as the spectra over different time periods. The source model includes all the point sources from the second {\it Fermi}-LAT catalog \citep[2FGL;][]{2012ApJS..199...31N} that lies within the 15$^{\circ}$ region of interest (ROI) of every source. Many sources are detected by {\it Fermi}-LAT after the release of the 2FGL catalog \citep[see e.g.][]{2011ATel.3579....1G} and thus not included in it. However, if these newly detected sources lie in the vicinity of a source of interest, they could have an effect on the results of the analysis. We therefore, search for possible new sources within the ROI of the five $\gamma$-NLSy1 galaxies studied here, by generating their residual TS map, using the tool {\tt gttsmap}. These residual TS maps are shown in Figure~\ref{fig:TSMAPS}. Using the criterion TS $> 25$, we find two un-modeled sources associated with each of the residual TS maps of 1H 0323+342 (RA, Dec. = 48$^{\circ}$.236, 36$^{\circ}$.260 and 56$^{\circ}$.092, 32$^{\circ}$.804; J2000) and PKS 2004$-$447 (RA, Dec. = 307$^{\circ}$.257, $-$42$^{\circ}$.592 and 307$^{\circ}$.505, $-$41$^{\circ}$.542; J2000) and one source each in the residual TS maps of SBS 0846+513 (RA, Dec. = 133$^{\circ}$.112, 55$^{\circ}$.497; J2000) and PMN J0948+0022 (RA, Dec. = 142$^{\circ}$.854, 0$^{\circ}$.497; J2000). No new source with TS $> 25$ is seen in the residual TS map of PKS 1502+036. These new sources are then modeled by a power law (PL) and are included in our analysis. However, we note that inclusion of these new sources have negligible effect on the results obtained without including them as they are faint. Further, both PL and log parabola (LP) models are used to parameterize the $\gamma$-ray spectra of the sources. For objects that fall within 7$^{\circ}$ of each of the target sources studied here, all parameters except the scaling factor are allowed to vary, for the sources that lie between 7$^{\circ}$ to 14$^{\circ}$ only the normalization factor is allowed to vary freely and for the remaining sources that lie outside 14$^{\circ}$, all parameters are fixed to the values given in the 2FGL catalog. We perform a first run of the ML analysis and then remove all sources from the model with TS $<$ 25. This updated model is then used to carry out a second ML analysis. If the likelihood fitting shows non-convergence, all parameters of the sources lying outside 7$^{\circ}$ from the center of the ROI are fixed to the values obtained from the average likelihood fitting and the fitting is repeated. In cases where we find the fitting to show non-convergence, we repeat the fit by fixing the photon indices of the sources further 1$^{\circ}$ inside from the edge of the ROI and this process is carried out iteratively till the analysis converges. To generate weekly binned light curves, we use a PL model for all $\gamma$-NLSy1 galaxies, as the PL indices obtained from this model show smaller statistical uncertainties when compared to those obtained from fits using more complex models. To generate daily averaged as well as six hour binned light curves, only the normalization parameter of the source of interest is kept free and the rest of the parameters are fixed to the values obtained from the five years averaged fitting. In all the cases, i.e. both for light curve as well as spectrum generation, we consider the source to be detected at the $\sim$ 3$\sigma$ level \citep[e.g.][]{1991ApJ...374..344K} if the condition TS $>$ 9 is satisfied in the respective time or energy bin. We calculate 2$\sigma$ upper limits whenever $\bigtriangleup~F_{\gamma}/F_{\gamma} > 0.5$ and/or 1 $<$ TS $<$ 9, where $\bigtriangleup~F_{\gamma}$ is the 1$\sigma$ error estimate in the $\gamma$-ray flux $F_{\gamma}$. This is found by varying the flux of the source given by {\tt gtlike}, till the TS reaches a value of 4 \citep[e.g.][]{2010ApJS..188..405A}. The upper limits are not calculated for TS $<$ 1. Primarily governed by the uncertainty in the effective area, the measured fluxes have energy dependent systematic uncertainties of around 10\% below 100 MeV, 5\% between 316 MeV to 10 GeV and 10\% above 10 GeV\footnote{http://fermi.gsfc.nasa.gov/ssc/data/analysis/LAT\_caveats.html}. Unless otherwise specified, all the errors quoted in this paper are 1$\sigma$ statistical uncertainties. \section {Results and Discussion}\label{result} \subsection{Long Term Flux Variability}\label{long_var} The $\gamma$-ray light curves of the five $\gamma$-NLSy1 galaxies, obtained using the procedures outlined in Section~\ref{data_red}, are shown in Figure~\ref{lc}. Depending on the $\gamma$-ray brightness, different time bins are used for different objects. For clarity, we have not shown upper limits. Variability in the light curve is tested by calculating fractional rms variability ($F_{\rm var}$) and variability probability by means of a simple $\chi^{2}$ test, using the recipes of \citet{2003MNRAS.345.1271V} and \citet{2010ApJ...722..520A} respectively and the results are shown in Table~\ref{statistic}. Bi-weekly binned light curve of 1H 0323+342 shows that it is nearly steady in $\gamma$-rays during the initial four years of the LAT observations, however, displays flaring behavior at two epochs in 2013, one around MJD 56,300 and the other around MJD 56,500. They are marked as H2 and H3 in Figure~\ref{lc}. The $F_{\rm var}$ for this source is found to be 0.498~$\pm$~0.046, which is on the higher side of that obtained for FSRQs ($\sim 0.24 \pm 0.01$) by \citet{2010ApJ...722..520A}, thus indicating substantial flux variations. The probability that the source has shown variation is $>$~99\% (see Table~\ref{statistic}). The average $\gamma$-ray flux during most of its quiescence is found to be $\sim$ 8 $\times$ 10$^{-8}$ ph cm$^{-2}$ s$^{-1}$. A maximum $\gamma$-ray flux of (5.11 $\pm$ 0.62) $\times$ 10$^{-7}$ ph cm$^{-2}$ s$^{-1}$~is observed in the bin centered at MJD 56,537 when a GeV flare was detected by {\it Fermi}-LAT \citep{2013ATel.5344....1C,2014ApJ...789..143P}. Five years averaged TS value of this source is found to be $\sim$ 723 which corresponds to $\sim$ 27$\sigma$ detection. SBS 0846+513 was in quiescence during the first two years of {\it Fermi} operation and thus was not included in the 2FGL catalog. This source was discovered in the $\gamma$-ray band only when {\it Fermi}-LAT detected a GeV flare during 2011 June \citep{2011ATel.3452....1D}. After that, many flaring activities have been observed from this source and they are shown by S1, S2, and S3 in Figure~\ref{lc}. A flux of $\sim$ 4 $\times$ 10$^{-8}$ ph cm$^{-2}$ s$^{-1}$~and TS $\sim$ 1700 is obtained from the average analysis. The probability that this source is variable is $>$~99\% and shows a value of 0.467~$\pm$~0.042 to the $F_{\rm var}$. PMN J0948+0022 is the first RL-NLSy1 galaxy detected in the $\gamma$-ray band by {\it Fermi}-LAT. An average analysis of $\sim$ 5 years of the LAT data gives the $\gamma$-ray flux and TS of $\sim$ 1.3 $\times$ 10$^{-7}$ ph cm$^{-2}$ s$^{-1}$~and $\sim$ 3300 respectively. From its weekly binned light curve, the peak flux value of (7.74~$\pm$~1.07)~$\times$~10$^{-7}$ ph cm$^{-2}$ s$^{-1}$~is observed during a recent GeV flare \citep{2013ATel.4694....1D}. The ratio of the maximum to minimum flux is obtained as $\sim$17 which is significantly higher than that noticed by \citet{2012A&A...548A.106F}. The primary reason of this difference is due to the time ranges covered by both the analyses. The peak flux is observed in the bin centered at MJD 56,297, which is not covered by \citet{2012A&A...548A.106F}. Few significant flaring activities are observed and we consider two of them for further analysis. These flaring periods are noted as P2 and P3 in Figure~\ref{lc}. We find a value of $F_{\rm var}$ of 0.443~$\pm$~0.029. This clearly hints at the existence of substantial flux variability in this source. PKS 1502+036 has been detected by {\it Fermi}-LAT most of the times, however, at a low flux level of $\sim$ 4.5 $\times$ 10$^{-8}$ ph cm$^{-2}$ s$^{-1}$, as can be seen in its monthly binned light curve in Figure~\ref{lc}. The five years averaged analysis gives a TS value of $\sim$ 435 which corresponds to $\sim$~20$\sigma$ detection. On comparing the results obtained here with that of \citet{2013MNRAS.433..952D}, the average flux and photon index values are found to be quite similar despite the fact that we have included one more year of the LAT data as compared to theirs. The similarity seen in both analyses can be justified by the fact that PKS 1502+036 is hardly detected by {\it Fermi}-LAT in the period MJD 56,235$-$56,597 (see the third panel from top in Figure~\ref{lc}) which is not covered by them. The $\chi^2$ analysis classifies the source to be a non-variable. PKS 2004$-$447 has the lowest values for the $\gamma$-ray flux (1.42~$\pm$~0.24~$\times$~10$^{-8}$~ph cm$^{-2}$ s$^{-1}$) and the TS (106) among $\gamma$-NLSy1 galaxies and was detected by {\it Fermi}-LAT only on a few occasions. Hence no definite conclusion can be drawn about its flux variability behavior owing to the low photon statistics as we could not get a significant value for the $F_{\rm var}$. \subsection{Short Term Flux Variability and Jet Energetics}\label{small_var} Blazars are known to vary on sub-hour time scales \citep[e.g.,][]{2011A&A...530A..77F,2013ApJ...766L..11S} and this could constrain the size as well as the location of the $\gamma$-ray emitting region \citep[see e.g.,][]{2010MNRAS.405L..94T,2013MNRAS.431..824B}. We, therefore, search for the presence of flux variability over short time scales (of the order of hours) in the light curves of the $\gamma$-NLSy1 galaxies. From the long term light curves of 1H 0323+342, SBS 0846+513, and PMN J0948+0022, we identify the epochs of high activity (shown in Figure~\ref{lc}) and generate one day as well as six hour binned light curves. We then scan the data to calculate the shortest flux doubling/halving time scale using the following equation: \begin{equation*} F(t)=F(t_{0}).2^{(t-t_{0})/\tau} \end{equation*} where $F(t)$ and $F(t_0)$ are the fluxes at time $t$ and $t_0$, respectively, and $\tau$ is the characteristic doubling/halving time scale. We also set the condition that the difference in flux at the epochs $t$ and $t_0$ is at least significant at 3$\sigma$ level \citep{2011A&A...530A..77F}. The recent GeV flare detected from 1H 0323+342 \citep{2013ATel.5344....1C} had its daily $\gamma$-ray flux as high as (1.75 $\pm$ 0.33) $\times$ 10$^{-6}$ ph cm$^{-2}$ s$^{-1}$~which is $\sim$~23 times larger compared to the five years averaged flux. The photon statistics during this epoch (2013 August 28 to 2013 September 1) is good enough to generate six hour binned light curve (see Figure~\ref{3hr_6hr}). In contrast to the earlier observations where $\gamma$-ray variability of the $\gamma$-NLSy1 galaxies were characterized over longer time scales \citep[$\geqslant$ 1 day; e.g.,][]{2011MNRAS.413.2365C,2012A&A...548A.106F}, this epoch of high activity provides evidence of extremely fast variability observed from a $\gamma$-NLSy1 galaxy. A flux doubling time as small as 3.09~$\pm$~0.85 hours (assuming exponential rise of the flare) is the fastest $\gamma$-ray variability ever observed from this class of AGN. In the time bin when the highest flux is measured, the TS value is 136 associated with a registered count of 34. The flux doubling time obtained here is significantly different from the value of 3.3~$\pm$~2.5 days reported by \citet{2011MNRAS.413.2365C} using the method of e-folding time scale. The primary reason for this contrast is the fact that 1H 0323+342 was in quiescence for most of the first four years of {\it Fermi} operation and hence could not produce enough photon statistics for studying short time scale variability. These findings are similar to that recently reported by \citet{2014ApJ...789..143P}. The highest flux in the six hour binned light curve is measured as (7.80 $\pm$ 1.43) $\times$ 10$^{-6}$ ph cm$^{-2}$ s$^{-1}$, which corresponds to an isotropic $\gamma$-ray luminosity of 2.82 $\times$ 10$^{46}$ erg\,s$^{-1}$, which is $\sim$~100 times higher than the five years averaged value. Total power radiated in the $\gamma$-ray band \citep{1997ApJ...484..108S}, thus, would be $L_{\gamma,em} \simeq L_{\gamma,iso}/2\Gamma^2 \simeq$ 2.20 $\times$ 10$^{44}$ erg\,s$^{-1}$~\citep[assuming a bulk Lorentz factor $\Gamma$=8;][]{2014ApJ...789..143P}, which is approximately 16\% of the Eddington luminosity considering a black hole mass of 10$^{7}$ \ensuremath{M_{\odot}}~\citep{2006ApJS..166..128Z}. Detection of a large flare in 2011 June led to the discovery of SBS 0846+513 in the $\gamma$-ray band. During this period of high activity, we find a maximum daily averaged flux of (8.20 $\pm$ 0.08) $\times$ 10$^{-7}$ ph cm$^{-2}$ s$^{-1}$, which is almost 20 times higher than its five years averaged flux. When the data are further re-binned using six hour bins, we notice a high $\gamma$-ray flux of (9.37 $\pm$ 1.74) $\times$ 10$^{-7}$ ph cm$^{-2}$ s$^{-1}$. The apparent isotropic $\gamma$-ray luminosity is found to be $\sim$ 1.44 $\times$ 10$^{48}$ erg\,s$^{-1}$. Blazars, in general, and FSRQs in particular, are known to be emitters of such high power in the $\gamma$-ray band. If a bulk Lorentz factor of 15 \citep{2012MNRAS.426..317D} is assumed, we estimate that the total radiated power in $\gamma$-rays would be $L_{\gamma,em}~\simeq$ 3.21 $\times$ 10$^{45}$ erg\,s$^{-1}$~which is about 93\% of the Eddington luminosity assuming a BH mass of 10$^{7.4}$ \ensuremath{M_{\odot}}~\citep{2008ApJ...685..801Y}, a value generally found for powerful FSRQs \citep[e.g.][]{2012Sci...338.1445N}. We find many instances during this flaring period when the apparent isotropic $\gamma$-ray luminosity exceeds 10$^{48}$ erg\,s$^{-1}$~(e.g., MJD 55,742 and 55,745). During the GeV flare in 2011 June, we find the shortest flux halving time scale of $\sim$ 25.6~$\pm$~11.0 hours. PMN J0948+0022 is the most $\gamma$-ray luminous NLSy1 galaxy ever detected by {\it Fermi}-LAT. During its recent GeV flare \citep{2013ATel.4694....1D} we observe a maximum one day averaged flux of (1.76 $\pm$ 0.32) $\times$ 10$^{-6}$ ph cm$^{-2}$ s$^{-1}$~on MJD 56,293 which is about 14 times the five years averaged value and the highest daily binned flux observed from the source in our analysis. This highest one day binned flux is similar to that noted by \citet{2012A&A...548A.106F}. On further dividing this active period into six hour bins, we find a maximum flux value of (2.40 $\pm$ 0.83) $\times$ 10$^{-6}$ ph cm$^{-2}$ s$^{-1}$. This corresponds to an isotropic $\gamma$-ray luminosity of 1.87 $\times$ 10$^{48}$ erg\,s$^{-1}$~which is $\sim$ 20 times its five years averaged value. Total radiated power in the $\gamma$-ray band is $L_{\gamma,em}~\simeq 4.16 \times 10^{45}$ erg\,s$^{-1}$~(assuming a bulk Lorentz factor of $\Gamma = 15$). If its BH mass is assumed to be $10^{7.5} M_{\odot}$~\citep{2008ApJ...685..801Y}, we find that the radiated $\gamma$-ray power is $\sim$ 95\% of the Eddington luminosity. The shortest time scale of variability observed from this source is found to be 74.7~$\pm$~27.6 hours, which is comparable to that reported by \citet{2012A&A...548A.106F} and \citet{2011MNRAS.413.2365C}. \subsection{Gamma-ray Spectrum}\label{gamma_spec} Analysis of the $\gamma$-ray spectral shape is done using two spectral models: PL ({$dN/dE \propto E^{-\Gamma_{\gamma}}$}), where $\Gamma_{\gamma}$ is the photon index and LP ({ $dN/dE \propto (E/E_{o})^{-\alpha-\beta log({\it E/E_{o}})}$}) where $E_o$ is the reference energy which is fixed at 300 MeV, $\alpha$ is the photon index at $E_o$ and $\beta$ is the curvature index which defines the curvature around the peak. The low photon statistics of all the sources precludes us to analyze the $\gamma$-ray spectral shape using more complex models such as broken power law (BPL) or power law with exponential cut-off. The spectral analysis is performed using a binned ML estimator {\tt gtlike} following the prescription given in Section~\ref{data_red}. We use a likelihood ratio test \citep{1996ApJ...461..396M} to check the PL model (null hypothesis) against the LP model (alternative hypothesis). Following \citet{2012ApJS..199...31N}, the curvature of test statistic $TS_{\rm curve}$ = 2(log L$_{\rm LP}-$log L$_{\rm PL}$) is also calculated. Presence of a significant curvature is tested by setting the condition $TS_{\rm curve} > 16$. \subsubsection{Average Spectral Analysis}\label{avg_spec} The results of the average spectral analysis of the five $\gamma$-NLSy1 galaxies using five years of {\it Fermi}-LAT data are tabulated in Table~\ref{avg_res}. Two sources, 1H 0323+342 and PMN J0948+0022, are found to have a significant curvature in their $\gamma$-ray spectrum. The value of $TS_{\rm curve}$ reported in Table~\ref{avg_res} shows that the LP model describes their spectrum better than the PL model. A $TS_{\rm curve}$ value of $\sim$~15 for SBS 0846+513 indicates a possible curvature in its $\gamma$-ray spectrum. For the remaining two sources, no significant curvature is noticed. This is possibly due to low photon statistics. However, in future, when more $\gamma$-ray data would be accumulated, it is likely that they too show a significant deviation from the PL model. To show the departure from a PL behavior and for better visualization of the spectral shape, we separately perform a binned likelihood analysis on appropriately chosen energy bins covering 0.1$-$300 GeV and generate the SEDs. The results are shown in Figure~\ref{avg_spec_fig}. Since there is no signal beyond 30 GeV, the $\gamma$-ray spectra are shown up to 30 GeV only. In this figure, the solid red line represents the PL model while the LP model is plotted by green dashed line. From the residual plot (lower panel) of the PL model fits to the sources, it is evident that the spectrum of 1H 0323+342 and PMN J0948+0022 clearly deviate from the PL behavior, confirming the $TS_{\rm curve}$ test. Also, hints of curvature can be seen in the residual plot of SBS 0846+513. PMN J0948+0022 has the most prominent curvature in its $\gamma$-ray spectrum ($\sim 8\sigma$) whereas $\sim 4\sigma$ significant curvature is noticed in the spectrum of 1H 0323+342. Further, the curvature index $\beta$ for PKS 2004$-$447 is found to be high and appears non-realistic as there is no significant curvature found in the $\gamma$-ray spectrum of this source. Thus, the unusual $\beta$ value may be indicating the poor photon statistics associated with the observation and hence seems unreliable. \subsubsection{Spectral Analysis of Low and High Activities}\label{flr_spec} Of the five $\gamma$-NLSy1 galaxies in our sample, three sources 1H 0323+342, SBS 0846+513, and PMN J0948+0022 are relatively bright and have shown GeV flaring activities \citep[e.g.,][]{2013ATel.5344....1C,2011ATel.3452....1D,2010ATel.2733....1D}, thus enabling us to probe their spectral behavior at different activity states. From the light curves, we identify periods when the sources are found to be in relatively low and high states. These periods are marked in Figure~\ref{lc}. A high activity state refers to the period of detection of high flux by the LAT, as already reported in the literature. A low activity state is selected as the time period when the source is in a relatively fainter state. We note that the low activity states may not represent the actual quiescence of the sources. The idea here is to compare the spectral shapes at different activity periods and selection of true quiescence would not serve the purpose due to poor photon statistics. Many GeV flares from PMN J0948+0022 have been reported \citep{2013ATel.4694....1D,2011ATel.3429....1D,2010ATel.2733....1D} and we select the two brightest flares (P2 and P3) for further analysis. The time duration along with the results of the likelihood analysis are presented in Table~\ref{flare_res}. For better visualization, we also show the spectra of the sources along with their best-fit model in Figure~\ref{H3_P9} and \ref{sbs}. A statistically significant ($\sim$ 4$\sigma$) curvature is noticed in the low activity state $\gamma$-ray spectrum of 1H 0323+342. Interestingly, no curvature is found in the high activity states of this source. A substantial curvature ($\sim 5\sigma$) is noticed in the first flaring state (S1) of SBS 0846+513 which is similar to that reported by \citet{2012MNRAS.426..317D}. However, no curvature is found in the other flaring state of this source (period S2; MJD 56,018$-$56,170), whereas for the same period \citet{2013MNRAS.436..191D} have reported the detection of curvature. A curvature is also noticed in the low activity state (P1) of PMN J0948+0022 but not during the GeV outbursts. This behavior is in contrast to that observed in bright blazars 3C 454.3 and PKS 1510$-$089 where a significant curvature is detected in their $\gamma$-ray spectra during the outbursts \citep{2011ApJ...733L..26A,2010ApJ...721.1425A}. \subsubsection{Spectral Evolution}\label{spec_evol} We investigate here the changes in the spectra in relation to the brightness of the $\gamma$-NLSy1 galaxies. Instead of attempting to provide a best description of a spectrum, we intend to search for possible trends in it. Hence, we evaluate the PL spectral indices during the time bins used for the light curves shown in Figure~\ref{lc}. The spectral index can be considered as a representative of the mean slope. The photon index of 1H 0323+342 is found to be harder in the high activity states. For example, in the bin centered at MJD 56,285 (peak flux in the interval H2 in Figure~\ref{lc}), the photon index value of 2.51~$\pm$~0.12 is significantly harder than the value of 2.78~$\pm$~0.05 found for the five year average. A hard photon index value of 2.49~$\pm$~0.11 is found during the 2013 August GeV outburst also. Similar behavior is also seen in SBS 0846+513. During its first GeV outburst, which led to the discovery of this source in the $\gamma$-ray band, the photon index is found to be 1.93$\pm$~0.09 which is typical of high synchrotron peaked blazars \citep{2011ApJ...743..171A}. Similar behavior of this source is earlier reported by \citet{2012MNRAS.426..317D}. The photon index of PMN J0948+0022 is also found to be harder during the high activity periods. During its first outburst of 2010 July, a photon index of 2.39~$\pm$~0.12 is obtained which is harder when compared to the five years averaged value of 2.62~$\pm$~0.02. This result is similar to that first reported by \citet{2011MNRAS.413.1671F}. A maximum flux of (1.21~$\pm$~0.36)~$\times$~10$^{-7}$~ph cm$^{-2}$ s$^{-1}$~is observed from PKS 1502+036 in the bin centered at MJD 55,477. The corresponding photon index during this period is 2.83~$\pm$~0.28 which is not significantly different from its five years averaged value of 2.63~$\pm$~0.05. Spectral evolution of PKS 2004$-$447 could not be ascertained owing to its faintness. Overall, though there are hints of spectral hardening for three out of the five $\gamma$-NLSy1 galaxies during flares, in comparison to their five year average behavior, no strong spectral change is noticed on shorter time scales. These results are in line with the findings of \citet{2010ApJ...710.1271A} and \citet{2012A&A...548A.106F}. To study the overall spectral behavior of these sources, we plot in Figure~\ref{ph_npred} the photon index against the flux, obtained from the light curve analysis. PKS 1502+036 shows a `softer when brighter' behavior whereas no conclusion can be drawn about PKS 2004$-$447 owing to the small number of data points. Similar results have been reported by \citet{2013ApJ...768...52P} also. Moreover, the remaining three sources seem to show a `softer when brighter' trend, up to a flux level of $\simeq 1.5 \times 10^{-7}$ ph cm$^{-2}$ s$^{-1}$. Above this flux value, these sources possibly hint for a `harder when brighter' trend. To statistically test these behaviors, we perform a Monte Carlo test that takes into account of the dispersion in flux and photon index measurements. For each observed pair of flux and index values, we re-sample it by extracting the data from a normal distribution centered on the observed value and standard deviation equal to the 1$\sigma$ error estimate. We do not perform this test on the data of PKS 1502+036 and PKS 2004$-$447 due to their small sample size. For lower fluxes (i.e., $F_{\gamma} < 1.5 \times 10^{-7}$ ph cm$^{-2}$ s$^{-1}$), the correlation coefficients ($\rho$) are found to be 0.21, 0.26, and 0.27 with 95\% confidence limits of $-0.27 \leqslant \rho \leqslant 0.61, -0.03 \leqslant \rho \leqslant 0.51$, and $-0.05 \leqslant \rho \leqslant 0.54$ respectively for 1H 0323+342, SBS 0846+513, and PMN J0948+0022. On the other hand, for higher fluxes ($F_{\gamma} > 1.5 \times 10^{-7}$ ph cm$^{-2}$ s$^{-1}$), the correlation coefficients are $-$0.32, $-$0.09, and $-$0.08 with 95\% confidence limits of $-0.67 \leqslant \rho \leqslant 0.14, -0.48 \leqslant \rho \leqslant 0.33$, and $-0.27 \leqslant \rho \leqslant 0.12$ respectively for 1H 0323+342, SBS 0846+513, and PMN J0948+0022. Clearly, based on the Monte Carlo analysis, it is difficult to claim for the presence of correlation between fluxes and photon indices. \subsection{Origin of Spectral Curvature/Break}\label{break_orign} The origin of the $\gamma$-ray spectral break detected by {\it Fermi}-LAT in many FSRQs is still an open question. Many theoretical models are available in the literature to explain the observed spectral break. One among the extrinsic causes could be due to the attenuation of $\gamma$-rays by photon-photon pair production on He II Lyman recombination lines within the BLR as explained by the double-absorber model of \citet{2010ApJ...717L.118P}. In this model, the $\gamma$-ray dissipation region should lie inside the BLR. However, by analyzing a sample of $\gamma$-ray bright blazars, \citet{2012ApJ...761....2H} have raised questions about the validity of the double-absorber model. The break energy predicted by this model is found to be inconsistent with observations. \citet{2010ApJ...714L.303F} proposed a scenario wherein the observed spectral break could be explained as a sum of hybrid scattering of accretion disk and BLR photons, but this solution requires a wind like profile for the BLR. Using an equipartition approach, \citet{2013ApJ...771L...4C} proposed a scenario in which the break in the GeV spectra occurs as a consequence of the Klein-Nishina (KN) effect on the Compton scattering of BLR photons by relativistic electrons having a curved particle distribution in the jet. Alternatively, the $\gamma$-ray spectral break can also result from intrinsic effects. In this intrinsic model, the break can happen if there is a cutoff in the energy distribution of particles that produce the $\gamma$-ray emission \citep{2009ApJ...699..817A}. Three out of the five $\gamma$-NLSy1 galaxies studied here, 1H 0323+342, SBS 0846+513 and PMN J0948+0022, show clear deviation from a simple PL model. Instead, the LP model seems to describe the data better. From the SED modeling of 1H 0323+342 and PMN J0948+0022, \citet{2009ApJ...707L.142A} constrained the $\gamma$-ray emission region in them to be well inside the BLR \citep[see also,][]{2012A&A...548A.106F,2014ApJ...789..143P}. In such a scenario, presence of the curvature in the $\gamma$-ray spectrum of these two sources can be explained on the basis of the KN effect. This is because the UV photons produced by the accretion disk and reprocessed by the BLR are blueshifted by a factor $\sim \Gamma$. Assuming a typical value of $\Gamma \simeq 10$, the approximate energy of these photons in the rest frame of the blob is about 10$^{16}$ Hz and hence the IC scattering with the electrons (having $\gamma \simeq 1000$) occurs under the mild KN regime. This hypothesis can also be tested by determining the energy of the highest energy photons detected by {\it Fermi}-LAT. The detection of very high energy photons ($>$ few tens of GeV) will clearly rule out the possibility of the emission region to be located inside the BLR. We used {\tt gtsrcprob} tool to determine the energy of the highest energy photon. This value is found to be 32.7 GeV (95.76\% detection probability) and 4.71 GeV (98.15\% detection probability) for 1H 0323+342 and PMN J0948+0022 respectively. The highest photon energy for 1H 0323+342 closely satisfies the BLR $\gamma$-ray transparency condition \citep{2009MNRAS.397..985G} and thus the origin of the $\gamma$-ray emission could lie inside the BLR. We note here that the non-detection of very high energy $\gamma$-ray photon does not ensure the emission region to be located inside the BLR but such a detection would have definitely ruled out this possibility. The $\gamma$-ray emission from SBS 0846+513 is well modeled by IC scattering of dusty torus photons \citep{2013MNRAS.436..191D} and this sets the location of the $\gamma$-ray emission region to be outside the BLR where IC scattering takes place under the Thomson regime. The most probable cause of the curvature/break in the $\gamma$-ray spectrum would be, thus, intrinsic to the emitting particle distribution. In this picture, the curvature must be visible in the $\gamma$-ray spectrum of SBS 0846+513 in all brightness states. However, the poor photon statistics hinders us to detect this curvature. Additionally, if the observed curvature is due to the curved particle energy distribution, a similar feature should be seen in the optical/UV part of the spectrum also, provided that the optical/UV part is dominated by the synchrotron radiation. Hence, a dedicated multi-wavelength study, similar to the one done for 3C 454.3 \citep{2013ApJ...771L...4C}, is desired to test these possible scenarios. \subsection{Gamma-ray loud NLSy1 Galaxies v/s Gamma-ray loud Blazars}\label{comparison} The $\gamma$-ray spectral curvature observed in some of the $\gamma$-NLSy1 galaxies studied here are commonly known in {\it Fermi}-LAT detected FSRQs \citep{2010ApJ...710.1271A}. Such breaks are also noticed in the LAT spectra of a few low and intermediate synchrotron peaked BL Lac objects. In these BL lac objects where $\gamma$-ray spectral break is observed, broad optical emission lines are reported in the quiescent states of few of them. The sources AO 0235+164 \citep{2007A&A...464..871R} and PKS 0537$-$441 \citep{2011MNRAS.414.2674G} are good examples for this. It is thus likely that these BL Lac objects could belong to the FSRQ class of AGN as pointed by \citet{2011MNRAS.414.2674G}. Moreover, there are observations of curvature in the X-ray spectra of many high frequency peaked BL Lac objects \citep[and references therein]{2011ApJ...739...73M}, but the presence of curvature in their $\gamma$-ray spectra is yet to be verified. Therefore, based on the available observations it is likely that the $\gamma$-ray spectral break is a characteristic feature of the FSRQ class of AGN. All the five sources studied here, have steep photon indices (Table~\ref{avg_res}) and the average value is found to be 2.55 $\pm$ 0.03. This is similar to the average photon index found for FSRQs (2.42 $\pm$ 0.17) in the 2LAC \citep{2011ApJ...743..171A} and is steeper than the value of 2.17 $\pm$ 0.12, 2.13 $\pm$ 0.14 and 1.90 $\pm$ 0.17 found for the low, intermediate and high synchrotron peaked BL Lac objects respectively. Thus, in terms of the average $\gamma$-ray photon index, these $\gamma$-NLSy1 galaxies are similar to FSRQs. However, the $\gamma$-ray luminosities of these five $\gamma$-NLSy1 galaxies are lower when compared to powerful FSRQs but higher than BL Lac objects \citep[e.g., Figure 6 of][]{2013ApJ...768...52P}. As some of the $\gamma$-NLSy1 galaxies studied here show evidences for the presence of spectral curvature and high flux variability, we argue based on the $\gamma$-ray properties, that these sources show resemblance to FSRQs. Positioning the FSRQs and BL Lac objects detected during the first three months of {\it Fermi} operation \citep{2009ApJ...700..597A} on the $\gamma$-ray luminosity ($L_\gamma$) v/s $\gamma$-ray spectral index ($\alpha_{\gamma}$) diagram, \citet{2009MNRAS.396L.105G} have postulated the existence of low BH mass FSRQs having steep $\alpha_{\gamma}$ and low $L_\gamma$. In the $L_{\gamma}$ v/s $\alpha_{\gamma}$ plane, the five $\gamma$-NLSy1 galaxies studied here tend to have steep $\alpha_{\gamma}$ values. Their $L_{\gamma}$ values are intermediate to the FSRQ 3C 454.3 and the BL Lac object Mrk 421 \citep{2013ApJ...768...52P}. Further, from the estimates of BH masses available in the literature, it is also found that these $\gamma$-NLSy1 galaxies host low mass BHs (Table~\ref{general}). They appear to be on the higher side of the BH masses known for NLSy1 galaxies in general \citep{2006AJ....132..531K,2008ApJ...685..801Y}, but lower than that known for powerful blazars \citep{2010MNRAS.402..497G}. It is thus likely that these $\gamma$-NLSy1 galaxies are the low BH mass FSRQs already predicted by \citet{2009MNRAS.396L.105G}. However, we note that the estimation of the BH mass of NLSy1 galaxies is still a matter of debate. For example, using the virial relationship \citet{2008ApJ...685..801Y} reported the central BH mass of PMN J0948+0022 as $\sim$~10$^{7.5}~M_{\odot}$, whereas using the SED modeling approach \citet{2009ApJ...707L.142A} found $\sim$~10$^{8.2}~M_{\odot}$. Recently \citet{2013MNRAS.431..210C} obtained the BH mass of this source as large as $\sim$~10$^{9.2}~M_{\odot}$. Further, on examination of the SEDs of these $\gamma$-NLSy1 galaxies, 1H 0323+342 \citep{2014ApJ...789..143P}, SBS 0846+513 \citep{2012MNRAS.426..317D}, PMN J0948+0022 \citep{2011MNRAS.413.1671F}, PKS 1502+036 and PKS 2004$-$447 \citep{2013ApJ...768...52P}, we find all of them to have Compton dominance (which is the ratio of inverse Compton to synchrotron peak luminosities in the SED) of the order of $\sim$ 10 similar to that known for powerful FSRQs. \section{Conclusions}\label{conclusion} In this paper, we present the detailed analysis of the $\gamma$-ray temporal and spectral behavior of the five RL-NLSy1 galaxies that are detected by {\it Fermi}-LAT with high significance. We summarize our main results as follows: \begin{enumerate} \item During the period 2008 August to 2013 November, three out of five $\gamma$-NLSy1 galaxies show significant flux variations. The same cannot be claimed for the remaining two sources due to their faintness and/or intrinsic low variability nature. The $F_{\rm var}$ values obtained are 0.498 $\pm$ 0.046, 0.467 $\pm$ 0.042 and 0.443 $\pm$ 0.029 for 1H 0323+342, SBS 0846+513, and PMN J0948+0022 respectively. These are larger than that reported for FSRQs ($0.24 \pm 0.01$) as well as BL Lac objects ($0.12 \pm 0.01, 0.07 \pm 0.02$, and $0.07 \pm 0.02$ for LSP, ISP, and HSP BL Lac objects respectively) by \citet{2010ApJ...722..520A}. This supports the idea that in terms of the $\gamma$-ray flux variations these sources are similar to the FSRQ class of AGN. More than one episode of flaring activities is seen in 1H 0323+342, SBS 0846+513 and PMN J0948+0022, whereas no such flaring activities is observed in PKS 1502+342 and PKS 2004$-$447. Results from Monte Carlo simulation, to test for the presence of a possible correlation between the fluxes and photon indices, do not support for any correlation both at lower as well as at higher flux levels. \item The observed average MeV$-$GeV $\gamma$-ray energy spectra of 1H 0323+342 and PMN J0948+0022 show significant deviation from a simple PL model. The data are well fit with the LP model. A hint for the presence of a curvature is found for SBS 0846+513. For PKS 1502+036 and PKS 2004$-$447, the PL model fits the observations very well. \item Three sources, 1H 0323+342, SBS 0846+513 and PMN J0948+0022 have shown flux variations in their $\gamma$-ray emission, thereby enabling us to study their spectral behavior in different brightness states. Again PL model do not well represent the observed spectra in some of their activity states. A statistically significant curvature is observed in the low, high and moderately active state $\gamma$-ray spectra of 1H 0323+342, SBS 0846+513 and PMN J0948+0022 respectively. \item The SED modeling of 1H 0323+342 and PMN J0948+0022 by \citet{2009ApJ...707L.142A} leads to the conclusion that the location of $\gamma$-ray emission in these sources is inside the BLR \citep[see also,][]{2012A&A...548A.106F,2014ApJ...789..143P}. It is thus likely that the observed $\gamma$-ray emission in these sources is due to IC scattering of BLR photons occurring under the KN regime. In such a scenario, a prominent curvature in the $\gamma$-ray spectra is expected \citep[see][for a similar argument]{2010ApJ...721.1425A}. For SBS 0846+513, the $\gamma$-ray emission is well modeled by IC scattering of dusty torus photons \citep{2013MNRAS.436..191D}. Since the production of $\gamma$-rays would be in the Thomson regime, the observed curvature could be intrinsic to the source and can be attributed to the break/curvature in the particle energy distribution. \item The average photon indices of the $\gamma$-NLSy1 galaxies, obtained from the PL fits, have values ranging between 2.2 and 2.8. This matches well with the average photon index (2.42 $\pm$ 0.17) found for FSRQs in the 2LAC. However these values are softer when compared to BL Lac objects \citep{2011ApJ...743..171A}. With the accumulation of more data by {\it Fermi} in the future and subsequent generation of high quality $\gamma$-ray spectra, it will be possible to constrain further various models proposed for the origin of MeV$-$GeV break and other $\gamma$-ray spectral properties of the $\gamma$-NLSy1 galaxies. Finally, as observed by \citet{2010ApJ...710.1271A} and \citet{2011ApJ...733...19T}, a break in the $\gamma$-ray spectrum is a characteristic feature of powerful FSRQs, which is also very clearly visible in the case of $\gamma$-NLSy1 galaxies. The origin of such curvature and break in their spectra could be because of different physical mechanisms taking place at different regions and at different scales. The overall observed properties of these $\gamma$-NLSy1 galaxies, indicate that they could be similar to powerful FSRQs but with low or moderate jet power \citep[see][for a comparison]{2009ApJ...707L.142A,2013MNRAS.436..191D,2013ApJ...768...52P}. \end{enumerate} \section*{Acknowledgments} We thank the referee for constructive comments that improved the presentation significantly. Use of {\it Hydra} cluster at the Indian Institute of Astrophysics is acknowledged. \bibliographystyle{apj}
\section{Introduction} Let~$\Omega \subset \mathbb R^2$ be a convex domain and let us denote by~$\gamma$ and~$dm_\gamma$ the standard Gaussian function and measure in~$\mathbb{R}^{2}$ respectively, that is \begin{equation*} \gamma(x,y):= \exp\left(-\frac{x^2+y^2}{2}\right) \qquad \mbox{and} \qquad dm_\gamma := \gamma(x,y) \, dx \, dy \,. \end{equation*} In this paper we consider the following Neumann eigenvalue problem for the Hermite operator \begin{equation}\label{problem} \left\{ \begin{array}{ll} - \mathop{\mathrm{div}}\nolimits (\gamma \nabla u) = \mu \gamma u & \mbox{in} \quad \Omega \,, \\ \\ \dfrac{\partial u}{\partial \bf{n}} =0 & \mbox{on} \quad \partial\Omega \,, \end{array} \right. \end{equation} where~$\bf{n}$ stands for the outward normal to $\partial \Omega$. As usual, we understand~\eqref{problem} as a spectral problem for the self-adjoint operator~$T$ in the Hilbert space $L^2_\gamma(\Omega) := L^2(\Omega,dm_\gamma)$ associated with the quadratic form $ t[u] := \|\nabla u\|^2 $, $ \mathsf{D}(t) := H_\gamma^1(\Omega) $. Here $\|\cdot\|$ denotes the norm in $L^2_\gamma(\Omega)$ and \begin{equation*} H_\gamma^1(\Omega) := \{u \in L^2_\gamma(\Omega) \ | \ \nabla u \in L^2_\gamma(\Omega) \} \end{equation*} is a weighted Sobolev space equipped with the norm $\sqrt{\|\cdot\|^2+\|\nabla \cdot\|^2}$. Since the embedding $H_\gamma^1(\Omega) \hookrightarrow L^2_\gamma(\Omega)$ is compact (see Remark \ref{oss} below), the spectrum of~$T$ is purely discrete. We arrange the eigenvalues of~$T$ in a non-decreasing sequence $\{\mu_n(\Omega)\}_{n=0}^{+\infty}$ where each eigenvalue is repeated according to its multiplicity. The first eigenfunction of~\eqref{problem} is clearly a constant with eigenvalue $\mu_{0}(\Omega)=0$ for any~$\Omega$. We shall be interested in the first non-trivial eigenvalue $\mu _{1}(\Omega)$ of~\eqref{problem}, which admits the following variational characterization \begin{equation}\label{minimax.intro} \mu _{1}(\Omega ) = \min \left\{ \dfrac{\displaystyle\int_{\Omega }|\nabla u|^{2}\,dm_\gamma} {\displaystyle\int_{\Omega } u^{2}\,dm_\gamma} \ : \ u\in H_\gamma^{1}(\Omega)\setminus \{0\} \,, \ \int_{\Omega } u \, dm_\gamma=0\right\} . \end{equation} A classical Poincar\'e-Wirtinger type inequality which goes back to Hermite (see for example \cite[Chapter II, p.~91 ff]{CH1}) states that \begin{equation}\label{poincare} \mu_1(\mathbb R^2)=1 \end{equation} and therefore \begin{equation*} \int_{\mathbb R^2 }\left( u-\int_{\mathbb R^2}u \, dm_\gamma\right) ^{2}dm_\gamma \leq \int_{\mathbb R^2}|\nabla u|^{2} \, dm_\gamma \,, \qquad \forall u\in H_\gamma^{1}(\mathbb R^2) \,. \end{equation*} Very recently an inequality analogous to \eqref{poincare} raised up in connection with the proof of the ``gap conjecture'' for bounded sets (see \cite{AC}). In \cite{Andrews-Ni_2012} the authors prove that if $\Omega$ is a bounded, convex set then \begin{equation}\label{an-ni} \mu_1(\Omega) \ge \mu_1\left(-\frac{\mathrm{d}(\Omega)}{2},\frac{\mathrm{d}(\Omega)}{2}\right) \end{equation} where $\mathrm{d}(\Omega)$ is the diameter of $\Omega$ and, here and throughout, $\mu_1\left(a,b\right)$ will denote the first nontrivial eigenvalue of the Sturm-Liouville problem \begin{equation}\label{N_1d} \left\{ \begin{array}{ll} -\left( \gamma_1v^{\prime }\right) ^{\prime }=\mu \gamma_1v & \mbox{in}\ \left( a,b\right) \,, \\ \\ v^{\prime }\left( a\right) =v^{\prime }\left( b\right) =0 \,, & \end{array} \right. \end{equation} with $-\infty \le a <b \le +\infty$ and \begin{equation*} \gamma_1(x):= \exp\left(-\frac{x^2}{2}\right). \end{equation*} Again, we understand~\eqref{N_1d} as a spectral problem for a self-adjoint operator with compact resolvent in $L^2_{\gamma_1}((a,b))$. It is well-known that \begin{equation}\label{l(R)} \mu _{1}(a,b)\geq 1 \quad \text{ with } \quad \mu _{1}(a,b)=1 \text{ \ if and only if \ }(a,b)=\mathbb{R}. \end{equation} As first result of this paper we extend the validity of~\eqref{an-ni} to any convex, possibly unbounded, planar domain (see \cite{BCHT} for the smooth case). \begin{theorem}\label{main copy(1)} Let $\Omega \subset \mathbb{R}^{2}$ be any convex domain. Then \begin{equation}\label{mi} \mu _{1}(\Omega )\geq \mu_1(\mathbb R)=1 \,. \end{equation} \end{theorem} The result is sharp in the sense that the equality in~\eqref{mi} is achieved for~$\Omega$ being any two-dimensional strip. It is natural to ask if the strips are the unique domains for which the equality in~\eqref{mi} is achieved. We provide a partial answer to the uniqueness question via the following theorem, which is the main result of this paper. \begin{theorem}\label{main} Let $\Omega $ be a convex subset of $ S_{y_{1},y_{2}} := \left\{ (x,y)\in \mathbb{R}^{2}:y_{1}<y<y_{2}\right\} $ for some $y_1,y_2\in \mathbb R$, $y_1 < y_2$. If $\mu _{1}(\Omega )=1$, then $\Omega $ is a strip. \end{theorem} Inequality \eqref{an-ni} is a Payne-Weinberger type inequality for the Hermite operator. We recall that the classical Payne-Weinberger inequality states that the first nontrivial eigenvalue of the Neumann Laplacian in a bounded convex set~$\Omega$, $\mu_1^{\Delta}(\Omega)$, satisfies the following bound \begin{equation}\label{PW} \mu_1^{{\Delta}}(\Omega)\ge \frac{\pi^2}{\mathrm{d}(\Omega)^2}, \end{equation} where $\pi^2/\mathrm{d}(\Omega)^{2}$ is the first nontrivial Neumann eigenvalue of the one-dimensional Laplacian in $\left( -\mathrm{d}(\Omega)/2,\mathrm{d}(\Omega)/2\right)$ (see \cite{PW}). The above estimate is the best bound that can be given in terms of the diameter alone in the sense that $\mu_1^{\Delta}(\Omega){\mathrm{d}(\Omega)^2}$ tends to $\pi^2$ for a parallepiped all but one of whose dimensions shrink to zero (see \cite{HW,V}). Estimate \eqref{mi} is sharp, not only asymptotically, since the equality sign is achieved when $\Omega$ is any strip~$S$. Indeed, it is straight-forward to verify that $\mu_1(S)=\mu_1(\mathbb R)=1$ for any strip~$S$. Hence the question faced in Theorem \ref{main} appears quite natural. The paper is organized as follows. Section~\ref{Sec.ext} contains the proof of Theorem~\ref{main copy(1)}, while Section~\ref{Sec.main} is devoted to the proof of Theorem~\ref{main}. The latter consists in various steps. We firstly deduce from~\eqref{mi} that any optimal set must be unbounded; then we show that it is possible to split an optimal set~$\Omega$ getting two sets that are still optimal and have Gaussian area $m_\gamma(\Omega)/2$. Repeating this procedure we obtain a sequence of thinner and thinner, optimal sets~$\Omega_k$ and we finally prove that there exists $a\in \overline \mathbb R$ such that $\mu_1(\Omega_k)$ converges as $k\to+\infty$ to $\mu_1(a,+\infty)$, which is strictly greater than 1 unless $a=-\infty$. This circumstance implies that $\Omega$ contains a straight-line, and hence~$\Omega$ is a strip. The convergence of $\mu_1(\Omega_k)$ to $\mu_1(a,+\infty)$ follows by a more general result established in Section~\ref{Sec.as}, where we actually prove a convergence of \emph{all} eigenvalues of~$T$ in thin domains to eigenvalues of a one-dimensional problem (see ~Theorem~\ref{Thm.convergence}). We also establish certain convergence of eigenfunctions. We believe that the convergence results are of independent interest, since our method of proof differs from known techniques in the case of the Neumann Laplacian in thin domains \cite{Arrieta_1995,Arrieta-Carvalho_2004,Jerison2,Schatzman_1996}. For optimisation results related to the present work, we refer the interested reader to \cite{BCT_p,ENT, FNT,BCM,CdB}. \section{Proof of Theorem~\ref{main copy(1)}}\label{Sec.ext} Repeating step by step the arguments contained in~\cite{BCHT}, Theorem~\ref{main copy(1)} is a consequence of the following extension result, which we believe is interesting on its own. \begin{theorem}\label{ext} Let $\Omega\subset \mathbb R^2$ be a convex domain and let $u \in H_\gamma^1(\Omega)$. Then there exists a function $\tilde u \in H_\gamma^1(\mathbb R^2)$ which is an extension of $u$ to $\mathbb R^2$, that is $$ \tilde u|_{\Omega}=u $$ and \begin{equation}\label{extension} \|\tilde u\|_{H_\gamma^1(\mathbb R^2)}\le C \, \|u\|_{H_\gamma^1(\Omega)} \,, \end{equation} where $C=C(\Omega)$. \end{theorem} \begin{proof} We preliminarily observe that, if $\Omega$ is bounded, the theorem can be immediately obtained from the classical result for the unweighted case (see for instance \cite[Thm.~4.4.1]{Evans-Gariepy}). So, from now on, we assume that $\Omega$ is unbounded and we adapt the arguments in \cite{Evans-Gariepy} to treat our case. We distinguish two cases: $0 \in \Omega$ and $0 \notin \Omega$. \medskip \fbox{\emph{Case 1:} $0 \in \Omega$.} The convexity of $\Omega$ ensures there exists a constant $L>0$ such that, for every $(x_0,y_0)\in \partial\Omega$, up to a rotation, there exist $r>0$ and an $L$-Lipschitz continuous function $\beta:\mathbb R \to [0,+\infty)$ such that, if we set $Q(x_0,y_0,r):=\{(x,y)\in \mathbb R^2:\> |x-x_0|<r,\> |y-y_0|<r\}$, it holds $$ \Omega \cap Q(x_0,y_0,r)=\{(x,y)\in \Omega: \> |x-x_0|<r,\> y_0-r<y<\beta(x)\}, \quad \max_{|x-x_0|<r}|\beta(x)-y_0|<\frac{r}{2}. $$ In other words, \begin{equation}\label{beta'} |\beta'(x)|\le L \qquad \mbox{for a.e.} \ \, x\in (x_0-r,x_0+r), \end{equation} with $L$ independent from $x_0,y_0,r$. Fix $(x_0,y_0)\in \partial\Omega$ and set $\Omega^i:=Q(x_0,y_0,r)\cap \Omega$ and $\Omega^e:=Q(x_0,y_0,r)\setminus \overline{\Omega}$. Let $u \in C^1(\overline{\Omega})$ and suppose for the moment that the support of $u$ is contained in $ Q(x_0,y_0,r)\cap \overline{\Omega}$. Set $$ u^e(x,y) := u(x,2\beta(x)-y) \qquad \mathrm{if} \> (x,y)\in \overline{\Omega^e}\,. $$ We get \begin{eqnarray*} \lefteqn{\int_{\Omega^e}{u^e(x,y)^2} \, \exp\left(-\frac{x^2+y^2}{2}\right)dxdy} \\ &&=\int_{\Omega^e}u(x,2\beta(x)-y)^2 \, \exp\left(-\frac{x^2+y^2}{2}\right) dxdy \\ &&= \int_{\Omega^i} u(s,t)^2 \, \exp\left(-\frac{s^2+(2\beta(s)-t)^2}{2}\right)dsdt \\ &&=\int_{\Omega^i} u(s,t)^2 \, \exp\left(-\frac{s^2+(2\beta(s)-t)^2}{2} +\frac{s^2+t^2}{2}\right) \, \exp\left(-\frac{s^2+t^2}{2}\right)dsdt. \end{eqnarray*} By elementary geometric considerations, taking into account the assumption $0\in \Omega$, it is easy to verify that \begin{equation} \exp\left(-\frac{s^2+(2\beta(s)-t)^2}{2}+\frac{s^2+t^2}{2}\right) \leq 1 \,, \qquad \forall \>(s,t)\in \Omega^i. \label{exp} \end{equation} Thus \begin{equation}\label{normau} \displaystyle\int_{\Omega^e}{u^e(x,y)}^{2} \, \exp\left(-\frac{x^2+y^2}{2}\right)dxdy \le \int_{\Omega^i} u(s,t)^2 \, \exp\left(-\frac{s^2+t^2}{2}\right)dsdt. \end{equation} On the other hand, by \eqref{beta'} and \eqref{exp} it holds \begin{eqnarray}\label{normaDu} \lefteqn{\int_{\Omega^e}|\nabla u^e(x,y)|^2 \,\exp\left(-\frac{x^2+y^2}{2}\right)dxdy} \notag \\ &&\le \int_{\Omega^i} \left[\left(\partial_s u(s,t)+2 \partial_t u(s,t) \beta'(s)\right)^2+\big(\partial_t u(s,t)\big)^2\right] \exp\left(-\frac{s^2+(2\beta(s)-t)^2}{2}\right)dsdt \notag \\ &&\le C(L)\int_{\Omega^i} |\nabla u(s,t)|^2 \,\exp\left(-\frac{s^2+t^2}{2}\right)dsdt. \end{eqnarray} Define $$ \tilde u:=\left\{ \begin{array}{ll} u & \mathrm{on}\quad \overline{\Omega^i} \,, \\ u^e & \mathrm{on}\quad \overline{\Omega^e} \,, \\ 0 & \mathrm{on} \quad \mathbb R^2 \setminus(\Omega^i\cup\Omega^e) \,. \end{array} \right. $$ If~$\Omega $ contains the origin and the support of $u$ is contained in $Q(x_0,y_0,r)\cap \overline{\Omega}$, then~\eqref{normau} and~\eqref{normaDu} imply~\eqref{extension} with $C=C(L)$. Now assume that $u \in C^1(\overline{\Omega})$ and drop the restriction on its support. Clearly $\partial \Omega$ is not compact, but we can cover $\partial \Omega$ with a countable family of squares $\{Q_{2k-1}\}=\{Q(x_0^k,y_0^k,r'_k)\}_{k=1}^\infty$, with $(x_0^k,y_0^k)\in\partial\Omega$ and $r'_k>0$ such that \eqref{beta'} holds true for each~$k$. Analogously, we can cover the set $\Omega\setminus \bigcup_{k=1}^{+\infty} Q_{2k-1}$ with a countable family of squares $\{Q_{2k}\}=\{Q(x_1^k,y_1^k,r''_k)\}_{k=1}^\infty$ with $(x_1^k,y_1^k)\in\Omega$ and $r''_k>0$. For the countable cover $\{Q_l\}_{l=1}^\infty$ of $\overline{\Omega}$ there is a partition of unity $\varphi_l$ subordinated to $Q_l$ with $\varphi_l$ smooth for each $l$ (see for instance~\cite[Thm. 3.14]{A}). Define $\widetilde{\varphi_l u}$ as above when~$l$ is odd and set $$ \tilde u := \sum_{l \>\mathrm{odd}}\widetilde{\varphi_l u} +\sum_{l \>\mathrm{even}} \varphi_l u. $$ Clearly $\tilde u$ satisfies \eqref{extension}. Finally if $u \in H_\gamma^1(\Omega)$, the claim follows by approximation arguments and the proof of Case~1 is accomplished. \medskip \fbox{\emph{Case 2:} $0 \not\in \Omega$.} Suppose now that $0\notin \Omega $ and denote $d_{0} := \mathop{\mathrm{dist}}\nolimits(0,\partial \Omega )$. Let us fix a vector $(\delta_1, \delta_2)$, with $\sqrt{\delta_1^2+\delta_2^2}>d_0$, in such a way that the translation $ \Phi: \mathbb R^2 \to \mathbb R^2: \{(x,y) \mapsto (x-\delta_1,y-\delta_2) \} $ maps $\Omega $ onto a set $\Phi(\Omega )$ containing the origin. Defining \begin{equation*} v(x,y) := u(x+\delta_1,y+\delta_2)\exp \left( - \frac{x\delta_1 }{2}-\frac{\delta_1 ^{2}}{4}\right)\exp \left( - \frac{y\delta_2 }{2}-\frac{\delta_2 ^{2}}{4}\right) , \end{equation*} for every $(x,y)\in \Phi(\Omega)$, we have \begin{equation*} \int_{\Omega }u^{2} \, dm_\gamma = \int_{\Phi(\Omega )} v^{2} \, dm_\gamma \,. \end{equation*} Since by construction $\Phi(\Omega )$ contains the origin, there exists a function $\tilde{v}\in H_\gamma^{1}(\mathbb{R}^{2})$ such that $\tilde{v}\left\vert _{\Phi(\Omega )}\right. =v$ and \begin{equation*} \|\tilde{v}\|_{H_\gamma^{1}(\mathbb{R}^{2})} \leq C(L) \, \|v\|_{H_\gamma^{1}(\Phi(\Omega ))} \,. \end{equation*} Letting \begin{equation*} \tilde{u}(x,y) :=\tilde{v}(x-\delta_1 ,y-\delta_2)\exp \left( \frac{x\delta_1 }{2}-\frac{\delta_1 ^{2}}{4}\right) \exp \left( \frac{y\delta_2 }{2}-\frac{\delta_2 ^{2}}{4}\right) , \end{equation*} we finally get that $\tilde u|_\Omega=u$ and \begin{equation*} \|\tilde{u}\|_{H_\gamma^{1}(\mathbb{R}^{2})} \leq C(L,d_{0}) \, \|u\|_{H_\gamma^{1}(\Omega)}. \end{equation*} This completes the proof of the theorem. \end{proof} \begin{remark}\label{oss} Using the fact that $H^{1}_\gamma(\mathbb{R}^{2})$ is compactly embedded into $L^{2}_\gamma(\mathbb{R}^{2})$ (see for example~\cite{D}) and the above extension theorem one can easily deduce the compact embedding of $H^{1}_\gamma(\Omega)$ into $L^{2}_\gamma(\Omega)$ (see also \cite{FP,BCHT}). Therefore, by the classical spectral theory on compact self-adjoint operators, $\mu _{1}(\Omega )$ satisfies the variational characterization~\eqref{minimax.intro}. \end{remark} \section{Proof of Theorem~\ref{main}}\label{Sec.main} The main ingredient in our proof of Theorem~\ref{main} is the following lemma, which tells us that cutting the optimiser of~\eqref{mi} in two convex, unbounded sets with equal Gaussian area, we again get two optimisers. \begin{lemma} Let $\Omega$ be a convex subset of $S_{y_1,y_2}$ with $\mu_1(\Omega)=1$. Let $\bar y\in (y_1,y_2)$ be such that the straight-line $\{y=\bar y\}$ divides $\Omega$ into two convex subsets with equal Gaussian area $m_\gamma(\Omega)$. Then $$ \mu_1\big(\Omega \cap \{y<\bar y\}\big) =\mu_1\big(\Omega \cap \{y>\bar y\}\big)=1. $$ \end{lemma} \begin{proof} Let $u$ be an eigenfunction of (\ref{problem}) corresponding to $\mu _{1}(\Omega ).$ By~\eqref{minimax.intro}, we know that $\int_{\Omega} u \, dm_\gamma=0$ and \begin{equation*} 1=\dfrac{\int_{\Omega }|\nabla u|^{2}\,dm_\gamma} {\int_{\Omega }u^{2}\,dm_\gamma} \,. \end{equation*} For each $\alpha \in [0,2\pi]$ there is a unique straight-line $r_\alpha$ orthogonal to $(\cos \alpha,\sin \alpha)$ such that it divides $\Omega$ into two convex sets $\Omega_\alpha',\Omega_\alpha''$ with equal Gaussian measure. Let $I(\alpha):=\int_{\Omega_\alpha'}u \, dm_\gamma$. Since $I(\alpha)=-I(\alpha+\pi)$, by continuity there is $\bar \alpha $ such that $I(\bar \alpha)=0$. Now we claim that $r_{\bar \alpha}$ is parallel to the $x$-axis. Note firstly that $\Omega_{\bar \alpha}'$ and $\Omega_{\bar \alpha}''$ are obviously convex and by \eqref{an-ni}, \eqref{l(R)} and \eqref{mi} we have \begin{equation} \mu _{1}(\Omega_{\bar \alpha}')\geq 1 \,, \qquad \mu_1(\Omega_{\bar \alpha}'')\ge 1 \,. \label{mu+->=1} \end{equation} Moreover, it is immediate to verify that \begin{equation*} 1=\mu_1(\Omega) =\dfrac{\int_{\Omega_{\bar \alpha}'}|\nabla u|^{2}\,dm_\gamma +\int_{\Omega_{\bar \alpha}''}|\nabla u|^{2}\,dm_\gamma} {\int_{\Omega_{\bar \alpha}'}u^{2}\,dm_\gamma +\int_{\Omega_{\bar \alpha}''}u^{2}\,dm_\gamma} \geq \min \left\{ \dfrac{\int_{\Omega_{\bar \alpha}'}|\nabla u|^{2}\,dm_\gamma} {\int_{\Omega_{\bar \alpha}'}u^{2}\,dm_\gamma}, \dfrac{\int_{\Omega_{\bar \alpha}''}|\nabla u|^{2}\,dm_\gamma} {\int_{\Omega_{\bar \alpha}''}u^{2}\,dm_\gamma}\right\}, \end{equation*} with equality holding if and only if $$ \dfrac{\int_{\Omega_{\bar \alpha}'}|\nabla u|^{2}\,dm_\gamma} {\int_{\Omega_{\bar \alpha}'}u^{2}\,dm_\gamma} =\dfrac{\int_{\Omega_{\bar \alpha}''}|\nabla u|^{2} \, dm_\gamma}{ \int_{\Omega_{\bar \alpha}''}u^{2}\,dm_\gamma}. $$ Without loss of generality we can assume that \begin{equation*} \min \left\{ \dfrac{\int_{\Omega_{\bar \alpha}'}|\nabla u|^{2}\,dm_\gamma} {\int_{\Omega_{\bar \alpha}'}u^{2}\,dm_\gamma}, \dfrac{\int_{\Omega_{\bar \alpha}''}|\nabla u|^{2}\,dm_\gamma} {\int_{\Omega_{\bar \alpha}''}u^{2}\,dm_\gamma}\right\} =\dfrac{\int_{\Omega_{\bar \alpha}'}|\nabla u|^{2}\,dm_\gamma} {\int_{\Omega_{\bar \alpha}'}u^{2}\,dm_\gamma}. \end{equation*} Finally, (\ref{mu+->=1}) ensures that \begin{equation} 1=\mu _{1}(\Omega )=\mu _{1}(\Omega_{\bar \alpha}') =\mu _{1}(\Omega_{\bar \alpha}'') \,. \label{1=m} \end{equation} Now we want to show that both $\Omega_{\bar \alpha}'$ and $\Omega_{\bar \alpha}''$ are unbounded, and hence $r_{\bar \alpha}$ is parallel to the $x$-axis. Suppose by contradiction that, for instance, $\Omega_{\bar \alpha}'$ is bounded. In such a case \eqref{an-ni} yields \begin{equation*} \mu _{1}(\Omega_{\bar \alpha}')\geq \mu _{1}\left( -\frac{\mathrm{d}(\Omega_{\bar \alpha}')}{2}, \frac{\mathrm{d}(\Omega_{\bar \alpha}')}{2}\right) . \end{equation*} Taking into account~(\ref{1=m}) and (\ref{l(R)}), we get that \begin{equation*} \mu_{1}\left( -\frac{\mathrm{d}(\Omega_{\bar \alpha}')}{2},\frac{\mathrm{d} (\Omega_{\bar \alpha}')}{2}\right) =1 \end{equation*} that is $\mathrm{d}(\Omega_{\bar \alpha}')=+\infty $, which is a contradiction. \end{proof} \begin{proof}[Proof of Theorem \ref{main}] By contradiction, let us assume that $\Omega \subset S_{y_1,y_2}$ is a convex domain different from a strip and $\mu_1(\Omega)=1$. Let us denote $$ \Omega=\left\{(x,y)\in \mathbb R^2: \, y_1<y<y_2,\, p(y)<x\right\}, $$ where $p$ is a convex, non-trivial function. From~\eqref{an-ni} and~\eqref{l(R)} it follows that~$\Omega$ is necessarily unbounded. By employing a separation of variables, we also deduce from~\eqref{an-ni} and~\eqref{l(R)} that~$\Omega$ cannot be a semi-strip. Finally, we may assume that $ \inf\{x: \exists\, y\in[y_1,y_2], \, (x,y)\in\Omega\} $ is finite (otherwise, we would have the finite supremum, which can be transferred to our situation by a reflection of the coordinate system). Repeating the procedure described in the above lemma, since at any step we are dividing into two convex subsets with equal Gaussian area, we can obtain a sequence of unbounded convex domains \begin{equation}\label{Omega.cut} \Omega_{\epsilon_k} := \left\{ (x,y)\in \mathbb{R}^{2}:\ y_0<y<d_k,p(y)<x \right\} = \left\{ (x,y)\in \Omega:\ y_0<y<d_k \right\} \end{equation} such that $$ \mu _{1}(\Omega_{\epsilon_k}) =1 \,, \qquad \epsilon_{k} := d_{k}-y_0 \xrightarrow[k \to +\infty]{} 0 \,. $$ Here the point~$y_0$ is chosen in such a way that $ p^{\prime}(y_0) \not= 0 $, which is always possible because the situation of semi-strips has been excluded. Without loss of generality (reflecting again the coordinate system if necessary), we may in fact assume \begin{equation}\label{Ass.phi} p^{\prime}(y_0) > 0 \,, \end{equation} so that~$\phi$ is increasing on $[y_0,d_k]$ whenever~$k$ is sufficiently large. Applying now a more general convergence result for eigenvalues in thin Neumann domains that we shall establish in the following section (Theorem~\ref{Thm.convergence}), we have \begin{lemma}\label{Lem.convergence} $ \displaystyle \lim_{k\to\infty}\mu_1(\Omega_{\epsilon_k}) =\mu_1(p^{-1}(y_0),+\infty) $. \end{lemma} \noindent Since $\mu _{1}(\Omega_{\epsilon_k})$ equals~$1$ for every~$k$, we conclude that $$ \mu_1\big(p^{-1}(y_0),+\infty\big) = 1 \,. $$ However, from \eqref{l(R)}, we then deduce that $p^{-1}(y_0)=-\infty$, which contradicts our assumptions from the beginning of the proof. In other words, $\Omega$ contains a straight-line and the theorem immediately follows. \end{proof} It thus remains to establish Lemma~\ref{Lem.convergence}. \section{Eigenvalue asymptotics in thin strips}\label{Sec.as} In this section we establish Lemma~\ref{Lem.convergence} as a consequence of a general result about convergence of \emph{all} eigenvalues of~$T$ in thin domains of the type~\eqref{Omega.cut}. \subsection{The geometric setting} Let $f:[0,+\infty) \to [0,+\infty)$ be a concave non-decreasing continuous non-trivial function such that $f(0) = 0$ (the case $f(0)>0$ is actually much easier to deal with). Given a positive number $\varepsilon < \sup f$, we put $$ f_\varepsilon(x) := \min\{\varepsilon,f(x)\} $$ and define an unbounded domain $$ \Omega_\varepsilon := \{(x,y)\in\mathbb{R}^2 : 0 < x \,, \ 0<y<f_\varepsilon(x)\} \,. $$ Clearly, \eqref{Omega.cut} can be cast into this form after identifying $f=p^{-1}$ and a translation. However, keeping in mind that the problem~\eqref{problem} is not translation-invariant, we accordingly change the definition of the Gaussian weight throughout this section $$ \gamma(x,y) := \exp\left(-\frac{(x_0+x)^2+(y_0+y)^2}{2}\right) \,. $$ Here~$y_0$ is primarily thought as the point from~\eqref{Omega.cut} and~$x_0$ is then such that $(x_0,y_0) \in \Omega_{\epsilon_k}$. For the results established in this section, however, $x_0$ and $y_0$ can be thought as arbitrary real numbers. For our method to work, it is only important to assume~\eqref{Ass.phi}, which accordingly transfers to \begin{equation}\label{Ass} f'(0) < +\infty \,. \end{equation} \subsection{The analytic setting and main result} Keeping the translation we have made in mind, instead of~\eqref{problem} we equivalently consider the eigenvalue problem \begin{equation}\label{bv} \left\{ \begin{array}{ll} - \mathop{\mathrm{div}}\nolimits (\gamma \nabla u) = \mu \gamma u & \mbox{in} \quad \Omega_\varepsilon \,, \\ \\ \dfrac{\partial u}{\partial \bf{n}} =0 & \mbox{on} \quad \partial\Omega_\varepsilon \,. \end{array} \right. \end{equation} We understand~\eqref{bv} as a spectral problem for the self-adjoint operator~$T_\varepsilon$ in the Hilbert space $L^2_\gamma(\Omega_\varepsilon)$ associated with the quadratic form $ t_\varepsilon[u] := \|\nabla u\|_\varepsilon^2 $, $ \mathsf{D}(t_\varepsilon) := H_\gamma^1(\Omega_\varepsilon) $. Here $\|\cdot\|_\varepsilon$ denotes the norm in $L^2_\gamma(\Omega_\varepsilon)$. We arrange the eigenvalues of~$T_\varepsilon$ in a non-decreasing sequence $\{\mu_n(\Omega_\varepsilon)\}_{n \in \mathbb{N}}$ where each eigenvalue is repeated according to its multiplicity. In this paper we adopt the convention $0 \in \mathbb{N}$. We are interested in the behaviour of the spectrum as $\varepsilon \to 0$, particularly $\mu_1(\Omega_\varepsilon)$ because of Lemma~\ref{Lem.convergence}. It is expectable that the eigenvalues will be determined in the limit $\varepsilon \to 0$ by the one-dimensional problem \begin{equation}\label{bv.1D} \left\{ \begin{array}{ll} - (\gamma_0 \, u')' = \nu \gamma_0 u & \mbox{in} \quad (0,+\infty) \,, \\ \\ u' (0)=0, & \end{array} \right. \end{equation} where $$ \gamma_0(x) := \gamma(x,0) = \exp\left(-\frac{(x_0+x)^2+y_0^2}{2}\right) \,. $$ Again, we understand~\eqref{bv.1D} as a spectral problem for the self-adjoint operator~$T_0$ in the Hilbert space $L^2_{\gamma_0}((0,+\infty))$ associated with the quadratic form $ t_0[u] := \|\nabla u\|_0^2 $, $ \mathsf{D}(t_0) := H_{\gamma_0}^1((0,+\infty)) $, where $\|\cdot\|_0$ denotes the norm in $L^2_{\gamma_0}((0,+\infty))$. As above, we arrange the eigenvalues of~$T_0$ in a non-decreasing sequence $\{\nu_n\}_{n \in \mathbb{N}}$ where each eigenvalue is repeated according to its multiplicity. By construction, for each $n\in\mathbb N$, $\nu_n$ coincides with the eigenvalue $\mu_n(x_0,+\infty)$ defined in \eqref{N_1d}. In this section we prove the following convergence result. \begin{theorem}\label{Thm.convergence} Let $f:[0,+\infty) \to [0,+\infty)$ be a concave non-decreasing continuous non-trivial function such that $f(0) = 0$. Assume in addition~\eqref{Ass}. Then $$ \forall n\in\mathbb{N} \,, \qquad \mu_n(\Omega_\varepsilon) \xrightarrow[\varepsilon \to 0]{} \nu_n \,. $$ \end{theorem} We shall also establish certain convergence of eigenfunctions of~$T_\varepsilon$ to eigenfunctions of~$T_0$. Clearly, Lemma~\ref{Lem.convergence} is the case $n=1$ of this general theorem. The rest of this section is devoted to a proof of Theorem~\ref{Thm.convergence}. \subsection{From the moving to a fixed domain} Our main strategy is to map~$\Omega_\varepsilon$ into a fixed strip~$\Omega$. We introduce a refined mapping in order to effectively deal with the singular situation $f(0)=0$. Let $$ a_\varepsilon := \inf f_\varepsilon^{-1}(\{\varepsilon\}) \,. $$ By the definition of~$f_\varepsilon$ and since~$f$ is non-decreasing, $a_\varepsilon \to 0$ as $\varepsilon \to 0$ and $f_\varepsilon(x)=\varepsilon$ for all $x>a_\varepsilon$. If $f(0)>0$, then there exists $\varepsilon_0>0$ such that $a_\varepsilon=0$ for all $\varepsilon \leq \varepsilon_0$. On the other hand, if $f(0)=0$, then $a_\varepsilon>0$ for all $\varepsilon>0$. The troublesome situation is the latter, to which we have restricted from the beginning. In this case, we introduce an auxiliary function $$ g_\varepsilon(s) := \begin{cases} a_\varepsilon s + a_\varepsilon &\mbox{if} \quad s \in [-1,0) \,, \\ s+a_\varepsilon &\mbox{if} \quad s \in [0,+\infty) \,. \end{cases} $$ Since we are interested in the limit $\varepsilon \to 0$, we may henceforth assume \begin{equation}\label{Ass.small} \varepsilon \leq 1 \qquad \mbox{and} \qquad a_\varepsilon \leq 1 \,. \end{equation} Define $\varepsilon$-independent sets $$ \Omega_-:=(-1,0)\times(0,1) \,, \qquad \Omega_+:=(0,+\infty)\times(0,1) \,, \qquad \Omega:=(-1,+\infty)\times(0,1) \,. $$ The mapping \begin{equation}\label{layer} \mathcal{L}_\varepsilon : \Omega \to \Omega_\varepsilon : \left\{ (s,t) \mapsto \mathcal{L}_\varepsilon(s,t) := \big( g_\varepsilon(s), f_\varepsilon(g_\varepsilon(s)) \, t \big) \right\} \end{equation} represents a $C^{0,1}$-diffeomorphism between~$\Omega$ and~$\Omega_\varepsilon$ ($f$~is differentiable almost everywhere, as it is supposed to be concave). In this way, {we obtain a convenient parameterisation of~$\Omega_\varepsilon$ via the coordinates $(s,t) \in \Omega$} whose Jacobian is \begin{equation}\label{Jacobian} j_\varepsilon(s,t) = g_\varepsilon'(s) f_\varepsilon(g_\varepsilon(s)) \,. \end{equation} Note that the Jacobian is independent of~$t$ and singular at $s=-1$. Now we reconsider~\eqref{bv} in~$\Omega$. With the notation $$ \gamma_\varepsilon(s,t) := (\gamma \circ \mathcal{L}_\varepsilon)(s,t) = \exp\left(-\frac{[x_0+g_\varepsilon(s)]^2 +[y_0+f_\varepsilon(g_\varepsilon(s))t]^2}{2}\right) \,, $$ introduce the unitary transform $$ U_\varepsilon: L^2_\gamma(\Omega_\varepsilon) \to L^2_{\gamma_\varepsilon j_\varepsilon/\varepsilon}(\Omega) : \left\{ u \mapsto \sqrt{\varepsilon} \, u \circ \mathcal{L}_\varepsilon \right\} \,. $$ Here, in addition to the change of variables~\eqref{layer}, we also make an irrelevant scaling transform (so that the renormalised Jacobian~$j_\varepsilon/\varepsilon$ is~$1$ in~$\Omega_+$). The operators $H_\varepsilon := U_\varepsilon T_\varepsilon U_\varepsilon^{-1}$ and~$T_\varepsilon$ are isospectral. By definition, $H_\varepsilon$~is associated with the quadratic form $h_\varepsilon[\psi] := t_\varepsilon[U_\varepsilon^{-1}\psi]$, $\mathsf{D}(h_\varepsilon) := U_\varepsilon \mathsf{D}(t_\varepsilon)$. \begin{proposition}\label{Prop.form} Assume~\eqref{Ass}. Then \begin{align} h_\varepsilon[\psi] &= \int_\Omega \left[ \left( \frac{\partial_s \psi}{g_\varepsilon'} - \frac{f_\varepsilon'\circ g_\varepsilon}{f_\varepsilon \circ g_\varepsilon} \, t \, \partial_t \psi \right)^2 + \frac{(\partial_t \psi)^2}{(f_\varepsilon\circ g_\varepsilon)^2} \right] \gamma_\varepsilon \, g_\varepsilon' \, \frac{f_\varepsilon \circ g_\varepsilon}{\varepsilon} \, ds \, dt \,, \qquad \label{form1} \\ \mathsf{D}(h_\varepsilon) &\subset H_{\gamma_\varepsilon j_\varepsilon/\varepsilon}^1(\Omega) \,. \label{form2} \end{align} \end{proposition} \noindent Here we have started to simplify the notation by suppressing arguments of the functions. \begin{proof} The space $ \mathcal{D}_\varepsilon := C_0^1(\mathbb{R}^2) \upharpoonright\Omega_\varepsilon $ is a core of~$t_\varepsilon$. The transformed space $ \mathcal{D} := U_\varepsilon \mathcal{D}_\varepsilon $ is a subset of $C_0^0(\mathbb{R}^2) \upharpoonright \Omega$ consisting of Lipschitz continuous functions on~$\Omega$ which belong to $C^1(\overline{\Omega_-}) \oplus C^1(\overline{\Omega_+})$ (we do not have $C^1$ globally, because~$g_\varepsilon$ and~$f_\varepsilon$ are not smooth). For any $\psi \in \mathcal{D}$, it is easy to check~\eqref{form1}; this formula extends to all~$\psi$ from the domain $$ \mathsf{D}(h_\varepsilon) = \overline{\mathcal{D}}^{\|\cdot\|_{h_\varepsilon}} \,, \qquad \|\cdot\|_{h_\varepsilon} := \sqrt{h_\varepsilon[\cdot] + \|\cdot\|^2} \,, $$ where $\|\cdot\|$ denotes the norm of $L^2_{\gamma_\varepsilon j_\varepsilon/\varepsilon}(\Omega)$. Let $\psi \in \mathcal{D}$. Using elementary estimates, we easily check \begin{equation}\label{bound.form} h_\varepsilon^-[\psi] \leq h_\varepsilon[\psi] \end{equation} where \begin{align*} h_\varepsilon^-[\psi] &:= \delta \int_{\Omega} \left( \frac{\partial_s \psi}{g_\varepsilon'} \right)^2 \gamma_\varepsilon \, g_\varepsilon' \, \frac{f_\varepsilon \circ g_\varepsilon}{\varepsilon} \, ds \, dt + \left( 1 - \frac{\delta}{1-\delta} \, \|f_\varepsilon'\|_\infty^2 \right) \int_{\Omega} \frac{(\partial_t \psi)^2}{(f_\varepsilon \circ g_\varepsilon)^2} \, \gamma_\varepsilon \, g_\varepsilon' \, \frac{f_\varepsilon \circ g_\varepsilon}{\varepsilon} \, ds \, dt \end{align*} with any $\delta \in (0,1)$. Note that~$f_\varepsilon'$ is bounded under the assumption~\eqref{Ass} and the concavity. For any $\varepsilon > 0$, we can choose~$\delta$ so small that $h_\varepsilon^-[\psi]$ is composed of a sum of two non-negative terms ($\delta$~can be made independent of~$\varepsilon$ if we restrict the latter to a fixed bounded interval, say $(0,1]$, see \eqref{Ass.small}, because $ \|f_\varepsilon'\|_{L^\infty((0,1))} \leq \|f'\|_{L^\infty((0,1))} $, but this assumption is not needed for the property we are proving). Using that~$g_\varepsilon'$ is bounded for any fixed~$\varepsilon$ and the estimate $f_\varepsilon \circ g_\varepsilon \leq \varepsilon$, we thus deduce from~\eqref{bound.form} that there is a positive constant~$c_{\varepsilon,\delta}$ (again, this constant can be made independent of~$\varepsilon$ if $\varepsilon \leq 1$) such that $$ c_{\varepsilon,\delta} \, \|\psi\|_{H_{\gamma_\varepsilon j_\varepsilon/\varepsilon}^1(\Omega)}^2 \leq \|\psi\|_{h_\varepsilon} \,. $$ This proves~\eqref{form2} because $\mathcal{D}$ is dense in $H_{\gamma_\varepsilon j_\varepsilon/\varepsilon}^1(\Omega)$. \end{proof} \subsection{The eigenvalue equation} Recall that we denote the eigenvalues of~$T_\varepsilon$ (and hence~$H_\varepsilon$) by $\mu_n(\Omega_\varepsilon)$ with $n \in \mathbb{N}$ ($=\{0,1,\dots\}$). The~$(n+1)^\mathrm{th}$ eigenvalue can be characterised by the Rayleigh-Ritz variational formula \begin{equation}\label{minimax.as} \mu_n(\Omega_\varepsilon) = \inf_{\stackrel{\dim \mathfrak{L}_n=n+1} {\mathfrak{L}_n \subset \mathsf{D}(h_\varepsilon)}} \sup_{\psi \in \mathfrak{L}_n} \frac{h_\varepsilon[\psi]}{\|\psi\|^2} \,. \end{equation} \begin{proposition}\label{Prop.bound} For any~$n \in \mathbb{N}$, there exists a positive constant~$C_n$ such that for all $\varepsilon \leq 1$, $$ \mu_n(\Omega_\varepsilon) \leq C_n \,. $$ \end{proposition} \begin{proof} Assuming $\varepsilon \leq 1$, we have the following two-sided $\varepsilon$- and $t$-independent bound \begin{equation}\label{2-bound} \gamma_-(s)\leq \gamma_\varepsilon(s,t) \leq \gamma_+(s) \end{equation} valid for every $(s,t) \in \Omega_+$ with \begin{equation*} \gamma_-(s) := \exp\left(-\frac{(|x_0|+s+1)^2 +(|y_0|+1)^2}{2}\right) , \qquad \gamma_+(s) := \exp\left(-\frac{(-|x_0|+s)^2-2|y_0|}{2}\right) . \end{equation*} Using in addition that $g_\varepsilon'=1$ and $f_\varepsilon\circ g_\varepsilon = \varepsilon$ in~$\Omega_+$, we obviously have $$ \forall \psi \in C_0^\infty((0,+\infty)) \otimes \{1\} \,, \qquad \frac{h_\varepsilon[\psi]}{\|\psi\|^2} \leq \frac{\displaystyle \int_{\Omega_+} (\partial_s \psi)^2 \, \gamma_+(s) \, ds \, dt} {\displaystyle \int_{\Omega_+} \psi^2 \, \gamma_-(s) \, ds \, dt} \,. $$ It then follows from~\eqref{minimax.as} that the inequality of the proposition holds with the numbers $$ C_n := \inf_{\stackrel{\dim \mathfrak{L}_n=n+1} {\mathfrak{L}_n \subset C_0^\infty((0,+\infty))}} \sup_{\psi \in \mathfrak{L}_n} \frac{\displaystyle \int_{0}^{+\infty} \psi'(s)^2 \, \gamma_+(s) \, ds} {\displaystyle \int_{0}^{+\infty} \psi(s)^2 \, \gamma_-(s) \, ds} \,, $$ which are actually eigenvalues of the one-dimensional operator $-\gamma_-^{-1} \partial_s \gamma_+ \partial_s$ in $L^2_{\gamma_-}((0,+\infty))$, subject to Dirichlet boundary conditions. \end{proof} Let us now fix $n \in \mathbb{N}$ and abbreviate the~$(n+1)^\mathrm{th}$ eigenvalue of~$H_\varepsilon$ by $\mu_\varepsilon := \mu_n(\Omega_\varepsilon)$. We denote an eigenfunction corresponding to~$\mu_\varepsilon$ by~$\psi_\varepsilon$ and normalise it to~$1$ in $L^2_{\gamma_\varepsilon j_\varepsilon/\varepsilon}(\Omega)$, \emph{i.e.}, \begin{equation}\label{norm} \|\psi_\varepsilon\|=1 \,. \end{equation} for every admissible $\varepsilon > 0$. The weak formulation of the eigenvalue equation $H_\varepsilon \psi_\varepsilon = \mu_\varepsilon \psi_\varepsilon$ reads \begin{equation}\label{weak} \forall \phi \in \mathsf{D}(h_\varepsilon) \,, \qquad h_\varepsilon(\phi,\psi_\varepsilon) = \mu_\varepsilon \, (\phi,\psi_\varepsilon) \,, \end{equation} where $(\cdot,\cdot)$ stands for the inner product in $L^2_{\gamma_\varepsilon j_\varepsilon/\varepsilon}(\Omega)$ {and $h_\varepsilon(\cdot,\cdot)$ denotes the sesquilinear form corresponding to~$h_\varepsilon[\cdot]$}, that is $ \forall \phi \in \mathsf{D}(h_\varepsilon) $ \begin{align} \int_\Omega \left[ \left( \frac{\partial_s \psi_{\varepsilon}}{g_\varepsilon'} - \frac{f_\varepsilon'\circ g_\varepsilon}{f_\varepsilon \circ g_\varepsilon} \, t \, \partial_t \psi_{\varepsilon} \right)\left( \frac{\partial_s \phi}{g_\varepsilon'} - \frac{f_\varepsilon'\circ g_\varepsilon}{f_\varepsilon \circ g_\varepsilon} \, t \, \partial_t \phi \right) + \frac{(\partial_t \psi_{\varepsilon})}{(f_\varepsilon\circ g_\varepsilon) } \frac{(\partial_t \phi)}{(f_\varepsilon\circ g_\varepsilon) } \right] \gamma_\varepsilon \, g_\varepsilon' \, \frac{f_\varepsilon \circ g_\varepsilon}{\varepsilon} \, ds \, dt \qquad \label{weform} \\ \notag = \mu_\varepsilon \int_\Omega \psi_{\varepsilon}\, \phi \, \gamma_\varepsilon \, g_\varepsilon' \, \frac{f_\varepsilon \circ g_\varepsilon}{\varepsilon} \, ds \, dt \,. \end{align} \subsection{What happens in \texorpdfstring{$\Omega_+$}{Omega+}} Using $|t| \leq 1$, we easily verify \begin{equation}\label{2-bound.bis} \forall (s,t) \in \Omega_+ \,, \qquad \gamma_\varepsilon(s,t) \geq \rho_\varepsilon(s) \gamma_0(s) \,, \end{equation} where the function $$ \rho_\varepsilon(s) := \exp\left( -\frac{a_\varepsilon^2+2|x_0| a_\varepsilon + \varepsilon^2+2|y_0|\varepsilon}{2} \right) \exp(-a_\varepsilon s) $$ is converging pointwise to~$1$ as $\varepsilon \to 0$. Choosing $\phi = \psi_\varepsilon$ as a test function in~\eqref{weak} and using~\eqref{2-bound.bis} together with Proposition~\ref{Prop.bound} and~\eqref{norm}, we obtain \begin{equation}\label{bound.plus} \int_{\Omega_+} (\partial_s\psi_\varepsilon)^2 \, \rho_\varepsilon \, \gamma_0 \, ds \, dt + \int_{\Omega_+} \frac{(\partial_t\psi_\varepsilon)^2}{\varepsilon^2} \, \rho_\varepsilon \, \gamma_0 \, ds \, dt \leq h_\varepsilon[\psi_\varepsilon] = \mu_\varepsilon \|\psi_\varepsilon\|^2 \leq C \,. \end{equation} Here and in the sequel, we denote by~$C$ a generic constant which is independent of~$\varepsilon$ and may change its value from line to line. Writing \begin{equation}\label{dec.plus} \psi_\varepsilon(s) = \varphi_\varepsilon(s) + \eta_\varepsilon(s,t) \,, \end{equation} where \begin{equation}\label{orthogonal} \int_0^1 \eta_\varepsilon(s,t) \, dt = 0 \qquad \mbox{for a.e.\ } s \in (0,+\infty) \,, \end{equation} we deduce from the second term on the left hand side of~\eqref{bound.plus} \begin{equation}\label{eta.plus} \pi^2 \int_{\Omega_+} \eta_\varepsilon^2 \, \rho_\varepsilon\, \gamma_0 \, ds \, dt \leq \int_{\Omega_+} (\partial_t\eta_\varepsilon)^2 \, \rho_\varepsilon \, \gamma_0 \, ds \, dt \leq C \varepsilon^2 \,. \end{equation} Differentiating~\eqref{orthogonal} with respect to~$s$, we may write $$ \int_{\Omega_+} (\partial_s\psi_\varepsilon)^2 \, \rho_\varepsilon \, \gamma_0 \, ds \, dt = \int_{\Omega_+} {\varphi_\varepsilon'}^2 \, \rho_\varepsilon \,\gamma_0 \, ds \, dt + \int_{\Omega_+} (\partial_s\eta_\varepsilon)^2 \, \rho_\varepsilon \, \gamma_0 \, ds \, dt $$ and putting this decomposition into~\eqref{bound.plus}, we get from the first term on the left hand side \begin{equation}\label{varphi'.plus} \int_0^{+\infty} {\varphi_\varepsilon'}^2 \, \rho_\varepsilon \, \gamma_0 \, ds \leq C \,, \qquad \int_0^{+\infty} (\partial_s\eta_\varepsilon)^2 \, \rho_\varepsilon \, \gamma_0 \, ds \leq C \,. \end{equation} At the same time, from~\eqref{norm} using~\eqref{2-bound.bis}, we obtain \begin{equation}\label{bound.plus.norm} \int_{\Omega_+} \varphi_\varepsilon^2 \, \rho_\varepsilon \, \gamma_0 \, ds \, dt +\int_{\Omega_+} \eta_\varepsilon^2 \, \rho_\varepsilon \, \gamma_0 \, ds \, dt =\int_{\Omega_+} \psi_\varepsilon^2 \, \rho_\varepsilon \, \gamma_0 \, ds \, dt \leq \|\psi_\varepsilon\|^2 = 1 \,, \end{equation} where the first equality employs~\eqref{orthogonal}. Consequently, \begin{equation}\label{varphi.plus} \int_0^{+\infty} \varphi_\varepsilon^2 \, \rho_\varepsilon \, \gamma_0 \, ds \leq 1 \,. \end{equation} Finally, employing the first inequality from~\eqref{varphi'.plus} and~\eqref{varphi.plus}, we get \begin{equation}\label{varphi'.plus.bis} \int_0^{+\infty} {(\sqrt{\rho_\varepsilon}\varphi_\varepsilon)'}^2 \, \gamma_0 \, ds \leq C \,. \end{equation} From~\eqref{varphi.plus} and~\eqref{varphi'.plus.bis}, we see that $\{\sqrt{\rho_\varepsilon}\varphi_\varepsilon\}_{\varepsilon>0}$ is a bounded family in $H^1_{\gamma_0}((0,+\infty))$ and therefore precompact in the weak topology of this space. Let~$\varphi_0$ be a weak limit point, \emph{i.e.}\ for a decreasing sequence of positive numbers $\{\varepsilon_i\}_{i \in \mathbb{N}}$ such that $\varepsilon_i \to 0$ as $i \to +\infty$, \begin{equation}\label{weak.limit} \sqrt{\rho_{\varepsilon_i}} \varphi_{\varepsilon_i} \xrightarrow[i\to+\infty]{w} \varphi_0 \qquad \mbox{in} \qquad H^1_{\gamma_0}((0,+\infty)) \,. \end{equation} Since $H^1_{\gamma_0}((0,+\infty))$ is compactly embedded in $L^2_{\gamma_0}((0,+\infty))$, we may assume \begin{equation}\label{strong.limit} \sqrt{\rho_{\varepsilon_i}}\varphi_{\varepsilon_i} \xrightarrow[i\to+\infty]{s} \varphi_0 \qquad \mbox{in} \qquad L^2_{\gamma_0}((0,+\infty)) \,. \end{equation} \subsection{What happens in \texorpdfstring{$\Omega_-$}{Omega-}} Here~$\gamma_\varepsilon$ can be estimated from below just by an $\varepsilon$-independent positive number, \emph{e.g.}, \begin{equation}\label{2-bound.minus} \forall (s,t) \in \Omega_- \,, \qquad \gamma_\varepsilon(s,t) \geq \exp\left( -\frac{(|x_0| +1)^2 + |y_0| +1)^2}{2} \right) \,. \end{equation} On the other hand, we need a lower bound to~$f_\varepsilon$. Employing that~$f$ is concave and non-decreasing, we can use \begin{equation}\label{f.bound} \forall s \in (-1,0) \,, \qquad f_\varepsilon(g_\varepsilon(s)) \geq \varepsilon \, (s+1) \,. \end{equation} Recall also that $g_\varepsilon'=a_\varepsilon$ on $(-1,0)$. Choosing $\phi = \psi_\varepsilon$ as a test function in~\eqref{weak} and using~\eqref{2-bound.minus} and~\eqref{f.bound}, we obtain \begin{equation}\label{bound.minus} \int_{\Omega_-} \left( \frac{\partial_s\psi_\varepsilon}{a_\varepsilon} - \frac{f_\varepsilon'\circ g_\varepsilon}{f_\varepsilon \circ g_\varepsilon} \, t \, \partial_t\psi_\varepsilon \right)^2 a_\varepsilon \, (s+1) \, ds \, dt + \int_{\Omega_-} \frac{(\partial_t\psi_\varepsilon)^2}{(f_\varepsilon \circ g_\varepsilon)^2} \, a_\varepsilon \, (s+1) \, ds \, dt \leq C \,. \end{equation} Assume~\eqref{Ass}. Using elementary estimates as in the proof of Proposition~\ref{Prop.form}, this inequality implies \begin{equation}\label{bound.minus.bis} \delta \int_{\Omega_-} \left( \frac{\partial_s\psi_\varepsilon}{a_\varepsilon} \right)^2 a_\varepsilon \, (s+1) \, ds \, dt + \left( 1 - \frac{\delta}{1-\delta} \, \|f_\varepsilon'\|_\infty^2 \right) \int_{\Omega_-} \frac{(\partial_t\psi_\varepsilon)^2}{(f_\varepsilon \circ g_\varepsilon)^2} \, a_\varepsilon \, (s+1) \, ds \, dt \leq C \end{equation} with any $\delta \in (0,1)$. We can choose~$\delta$ (independent of~$\varepsilon$ due to~\eqref{Ass.small}) so small that the left hand side of~\eqref{bound.minus.bis} is composed of a sum of two non-negative terms. Using in addition $f_\varepsilon \circ g_\varepsilon \leq \varepsilon$, we thus deduce from~\eqref{bound.minus.bis} \begin{equation*} \frac{1}{a_\varepsilon} \int_{\Omega_-} (\partial_s\psi_\varepsilon)^2 \, (s+1) \, ds \, dt + \frac{a_\varepsilon}{\varepsilon^2} \int_{\Omega_-} (\partial_t\psi_\varepsilon)^2 \, (s+1) \, ds \, dt \leq C \,. \end{equation*} Moreover, it follows from~\eqref{Ass} and the convexity bound \begin{equation}\label{convexity} \forall s \geq 0 \,, \qquad f(s) \leq f'(0) s \end{equation} that \begin{equation}\label{convexity.corol} \varepsilon \leq f'(0) \, a_\varepsilon \,. \end{equation} Hence \begin{equation}\label{bound.minus.bis2} \int_{\Omega_-} |{\nabla }\psi_\varepsilon|^2 \, (s+1) \, ds \, dt \leq C a_\varepsilon \,. \end{equation} Now we write ($\varphi_\varepsilon$ is constant!) \begin{equation}\label{dec.minus} \psi_\varepsilon(s,t) = \varphi_\varepsilon + \eta_\varepsilon(s,t) \,, \end{equation} where \begin{equation}\label{orthogonal.minus} \int_{\Omega_-} \eta_\varepsilon(s,t) \, (s+1) \, ds\, dt = 0 \,. \end{equation} Then we deduce from~\eqref{bound.minus.bis2} \begin{equation}\label{eta.minus} \pi^2 \int_{\Omega_-} \eta_\varepsilon^2 \, (s+1) \, ds \, dt \leq \int_{\Omega_-} |{\nabla }\eta_\varepsilon|^2 \, (s+1) \, ds \, dt \leq C a_\varepsilon \,. \end{equation} Note that~$\pi^2$ is indeed the minimum between the first non-zero Neumann eigenvalue in the interval of unit length and the first non-zero Neumann eigenvalue in the unit disk. At the same time, from~\eqref{norm} using~\eqref{2-bound.minus} and~\eqref{f.bound}, we obtain \begin{equation}\label{bound.minus.norm} \int_{\Omega_-} \varphi_\varepsilon^2 \, a_\varepsilon \, (s+1) \, ds \, dt +\int_{\Omega_-} \eta_\varepsilon^2 \, a_\varepsilon \, (s+1) \, ds \, dt = \int_{\Omega_-} \psi_\varepsilon^2 \, a_\varepsilon \, (s+1) \, ds \, dt \leq C \,, \end{equation} where the first equality employs~\eqref{orthogonal.minus}. Consequently, recalling that~$\varphi_\varepsilon$ is constant, \begin{equation}\label{varphi.minus} \varphi_\varepsilon^2 \, a_\varepsilon \leq C \qquad \mbox{on} \quad \Omega_- \,. \end{equation} \subsection{The limiting eigenvalue equation in \texorpdfstring{$\Omega_+$}{Omega+}} Now we consider~\eqref{weak} for the sequence $\{\varepsilon_i\}_{i\in\mathbb{N}}$ and a test function $\phi(s,t) = \varphi(s)$, where $\varphi \in C_0^\infty(\mathbb{R})$ is such that $\varphi'=0$ on $[-1,0]$, and take the limit $i \to +\infty$. We shall need a lower bound analogous to the upper bound~\eqref{convexity.corol}. From the fundamental theorem of calculus, we deduce \begin{equation}\label{Taylor} \forall s \in [0,a_\varepsilon] \,, \qquad f(s) \geq \big( \mathop{\mathrm{ess\;\!inf}}_{(0,a_\varepsilon)} f' \big) \, s \,. \end{equation} Note that the infimum cannot be zero unless~$f$ is trivial (we assume from the beginning $\varepsilon < \sup f$ and that~$f$ is non-decreasing) and that it converges to $f'(0)>0$ as $\varepsilon \to 0$. Consequently, for all sufficiently small~$\varepsilon$, we have \begin{equation}\label{convexity.corol.lower} \varepsilon \geq \frac{1}{2} \, f'(0) \, a_\varepsilon \,. \end{equation} At the same time, in analogy with~\eqref{2-bound.bis}, we have \begin{equation}\label{2-bound.new} \forall (s,t) \in \Omega_+ \,, \qquad \gamma_\varepsilon(s,t) \leq c_\varepsilon \rho_\varepsilon(s) \gamma_0(s) \,, \end{equation} where $$ c_\varepsilon := \exp\left( \frac{2 |x_0| a_\varepsilon + \varepsilon^2 + 2|y_0|\varepsilon}{2} \right) $$ is converging to~$1$ as $\varepsilon \to 0$. We first look at the right hand side of~\eqref{weak}. Using the decompositions~\eqref{dec.plus} and~\eqref{dec.minus}, we have \begin{multline*} (\varphi,\psi_{\varepsilon}) = \int_{\Omega_-} \varphi \, \varphi_{\varepsilon} \, \gamma_\varepsilon \, a_\varepsilon \, \frac{f_\varepsilon \circ g_\varepsilon}{\varepsilon} \, ds \, dt + \int_{\Omega_-} \varphi \, \eta_\varepsilon \, \gamma_\varepsilon \, a_\varepsilon \, \frac{f_\varepsilon \circ g_\varepsilon}{\varepsilon} \, ds \, dt \\ + \int_{\Omega_+} \varphi \, \varphi_\varepsilon \, \gamma_\varepsilon \, ds \, dt + \int_{\Omega_+} \varphi \, \eta_\varepsilon \, \gamma_\varepsilon \, ds \, dt \,. \end{multline*} Estimating $\gamma_\varepsilon \leq 1$ and using~\eqref{convexity} and~\eqref{convexity.corol.lower}, we get $$ \left| \int_{\Omega_-} \varphi \, \eta_\varepsilon \, \gamma_\varepsilon \, a_\varepsilon \, \frac{f_\varepsilon \circ g_\varepsilon}{\varepsilon} \, ds \, dt \right| \leq \frac{a_\varepsilon^2}{\varepsilon} f'(0) \int_{\Omega_-} |\varphi| \, |\eta_\varepsilon| \, (s+1) \, ds \, dt \leq 2 a_\varepsilon \int_{\Omega_-} |\varphi| \, |\eta_\varepsilon| \, (s+1) \, ds \, dt \,, $$ where the right hand side tends to zero as $\varepsilon \to 0$ due to the Schwarz inequality and~\eqref{eta.minus}. At the same time, recalling that~$\varphi_\varepsilon$ is constant in~$\Omega_-$, $$ \left| \int_{\Omega_-} \varphi \, \varphi_\varepsilon \, \gamma_\varepsilon \, a_\varepsilon \, \frac{f_\varepsilon \circ g_\varepsilon}{\varepsilon} \, ds \, dt \right| \leq \frac{a_\varepsilon^2}{\varepsilon} f'(0) \int_{\Omega_-} |\varphi| \, |\varphi_\varepsilon| \, (s+1) \, ds \, dt \leq 2 a_\varepsilon |\varphi_\varepsilon| \int_{\Omega_-} |\varphi| \, (s+1) \, ds \, dt \,, $$ where the right hand side tends to zero as $\varepsilon \to 0$ due to~\eqref{varphi.minus}. Using~\eqref{2-bound.new}, we also get $$ \left| \int_{\Omega_+} \varphi \, \eta_\varepsilon \, \gamma_\varepsilon \, ds \, dt \right| \leq c_\varepsilon \int_{\Omega_+} |\varphi| \, |\eta_\varepsilon| \, \rho_\varepsilon \, \gamma_0 \, ds \, dt \leq c_\varepsilon \sqrt{\int_{\Omega_+} \varphi^2 \, \gamma_0 \, ds \, dt} \sqrt{\int_{\Omega_+} \eta_\varepsilon^2 \, \rho_\varepsilon \, \gamma_0 \, ds \, dt} \,, $$ where the right hand side tends to zero as $\varepsilon \to 0$ due to~\eqref{eta.plus}. Finally, we write $$ \int_{\Omega_+} \varphi \, \varphi_{\varepsilon_i} \gamma_{\varepsilon_i} ds \, dt = \int_{\Omega_+} \varphi \, \varphi_{\varepsilon_i} \sqrt{\rho_{\varepsilon_i}} \gamma_{0} \, ds \, dt + \int_{\Omega_+} \varphi \, \varphi_{\varepsilon_i} \sqrt{\rho_{\varepsilon_i}} \gamma_{0} \, \left(\frac{\gamma_{\varepsilon_i}}{\sqrt{\rho_{\varepsilon_i}} \gamma_0} -1\right) \, ds \, dt \,. $$ Here the first term on the right hand side converges to $ \int_{\Omega_+} \varphi \, \varphi_{0} \, \gamma_{0} \, ds \, dt $ as $i \to +\infty$ due to~\eqref{weak.limit}, while the second term vanishes in the limit because of \begin{multline*} \left| \int_{\Omega_+} \varphi \, \varphi_{\varepsilon_i} \sqrt{\rho_{\varepsilon_i}} \gamma_{0} \, \left(\frac{\gamma_{\varepsilon_i}}{\sqrt{\rho_{\varepsilon_i}} \gamma_0} -1\right) \, ds \, dt \right| \\ \leq \sqrt{ \int_{\Omega_+} \varphi^2 \, \gamma_{0} \, \left(\frac{\gamma_{\varepsilon_i}}{\sqrt{\rho_{\varepsilon_i}} \gamma_0} -1\right)^2 \, ds \, dt } \sqrt{ \int_{\Omega_+} \varphi_{\varepsilon_i}^2 \rho_{\varepsilon_i} \gamma_{0} \, \, ds \, dt } \,. \end{multline*} Indeed the second term on the right hand side is bounded by~\eqref{bound.plus.norm}, while first term tends to zero as $i \to +\infty$ by the dominated convergence theorem. Summing up, \begin{equation}\label{lim.right} \lim_{i \to +\infty} (\varphi,\psi_{\varepsilon_i}) = \int_{0}^{+\infty} \varphi \, \varphi_{0} \, \gamma_{0} \, ds \,. \end{equation} Employing that the test function~$\varphi$ is constant on $[-1,0]$ and the decomposition~\eqref{dec.plus}, we have $$ h_\varepsilon(\varphi,\psi_{\varepsilon}) = \int_{\Omega_+} \varphi' \, \varphi_\varepsilon' \, \gamma_\varepsilon \, ds \, dt + \int_{\Omega_+} \varphi' \, \partial_s\eta_\varepsilon \, \gamma_\varepsilon \, ds \, dt \,. $$ Here the first term on the right hand side can treated in the same way as above with the conclusion $$ \int_{\Omega_+} \varphi' \, \varphi_{\varepsilon_i}' \, \gamma_{\varepsilon_i} \, ds \, dt \xrightarrow[i \to +\infty]{} \int_{\Omega_+} \varphi' \, \varphi_{0}' \, \gamma_{0} \, ds \, dt = \int_{0}^{+\infty} \varphi' \, \varphi_{0}' \, \gamma_{0} \, ds \,, $$ while we integrate by parts to handle the second term, $$ \int_{\Omega_+} \varphi' \, \partial_s\eta_\varepsilon \, \gamma_\varepsilon \, ds \, dt = - \int_{\Omega_+} \varphi'' \, \eta_\varepsilon \, \gamma_\varepsilon \, ds \, dt - \int_{\Omega_+} \varphi' \, \eta_\varepsilon \, \partial_s\gamma_\varepsilon \, ds \, dt \,. $$ Notice that the boundary terms vanish because~$\varphi$ has a compact support in~$\mathbb{R}$ and $\varphi'(0)=0$. As above, the first term on the right hand side vanishes as $\varepsilon \to 0$ due to~\eqref{eta.plus}. Similarly, $$ \begin{aligned} \left| \int_{\Omega_+} \varphi' \, \eta_\varepsilon \, \partial_s\gamma_\varepsilon \, ds \, dt \right| &\leq c_\varepsilon \int_{\Omega_+} |\varphi| \, |\eta_\varepsilon| \, \rho_\varepsilon \, \gamma_0 \, (x_0+s+a_\varepsilon) \, ds \, dt \\ &\leq c_\varepsilon \sqrt{\int_{\Omega_+} \varphi^2 \, \gamma_0 \, (x_0+s+a_\varepsilon)^2 \, ds \, dt} \sqrt{\int_{\Omega_+} \eta_\varepsilon^2 \, \rho_\varepsilon \, \gamma_0 \, ds \, dt} \,, \end{aligned} $$ where the right hand side tends to zero as $\varepsilon \to 0$ due to~\eqref{eta.plus}. Summing up, \begin{equation}\label{lim.left} \lim_{i \to +\infty} h_{\varepsilon_i}(\varphi,\psi_{\varepsilon_i}) = \int_{0}^{+\infty} \varphi' \, \varphi_{0}' \, \gamma_{0} \, ds \,. \end{equation} Since the set of functions $\varphi \in C_0^\infty(\mathbb{R})$ satisfying $\varphi'(0)=0$ is a core for the form domain of the operator~$T_0$, we conclude from~\eqref{lim.left} and~\eqref{lim.right} that~$\varphi_0$ belongs to~$\mathsf{D}(T_0)$ and solves the one-dimensional problems \begin{equation}\label{1Dproblems} \begin{aligned} T_0\varphi_0 &= \mu_0^+ \varphi_0 \,, \qquad & \mu_0^+ &:= \limsup_{i\to+\infty} \mu_{\varepsilon_i} \,, \\ T_0\varphi_0 &= \mu_0^- \varphi_0 \,, \qquad & \mu_0^- &:= \liminf_{i\to+\infty} \mu_{\varepsilon_i} \,. \end{aligned} \end{equation} If $\varphi_0 \not=0$ on $(0,+\infty)$, then~$\mu_0^\pm$ must coincide with some eigenvalues of~$T_0$. It remains to check that indeed $\varphi_0 \not=0$ on $(0,+\infty)$. \subsection{The limiting problem in \texorpdfstring{$\Omega_-$}{Omega-}: a crucial step} Define $$ \Omega_-' := (-1/2,0)\times(0,1) \,, \qquad \Omega_+' := (0,1/2)\times(0,1) \,, \qquad \Omega' := (-1/2,1/2)\times(0,1) \,. $$ From~\eqref{bound.minus.bis2} and~\eqref{bound.minus.norm}, we respectively have \begin{equation} \int_{\Omega_-'} |{\nabla }\psi_\varepsilon|^2 \, ds \, dt \leq 2C a_\varepsilon \,, \qquad \int_{\Omega_-'} \psi_\varepsilon^2 \, ds \, dt \leq \frac{2C}{a_\varepsilon} \,. \end{equation} At the same time, denoting $ m_0 := \min_{[0,1/2]} \gamma_0 $ and assuming $\varepsilon \leq 1$, from~\eqref{bound.plus} and~\eqref{bound.plus.norm}, we respectively get \begin{equation} \int_{\Omega_+'} |{\nabla }\psi_\varepsilon|^2 \, ds \, dt \leq \frac{C}{m_0 \, \rho_\varepsilon(1/2)} \,, \qquad \int_{\Omega_+'} \psi_\varepsilon^2 \, ds \, dt \leq \frac{1}{m_0 \, \rho_\varepsilon(1/2)} \,. \end{equation} Consequently, $\psi_\varepsilon \in H^1(\Omega')$ for any $\varepsilon \leq 1$ (although, in principle, $\|\psi_\varepsilon\|_{H^1(\Omega')}$ might not be uniformly bounded in~$\varepsilon$). It follows that the boundary values $\psi_\varepsilon(0-,t)$ and $\psi_\varepsilon(0+,t)$ exist in the sense of traces in~$\Omega_-'$ and~$\Omega_+'$, respectively, and they must be equal as functions of~$t$ in $L^2((0,1))$. Using the decompositions~\eqref{dec.plus} and~\eqref{dec.minus}, we therefore have, for almost every $t \in (0,1)$, \begin{multline*} [\varphi_\varepsilon(0-) - \varphi_\varepsilon(0+)]^2 = \left[ \eta_\varepsilon(0+,t) - \eta_\varepsilon(0-,t) \right]^2 \leq 2\,[\eta_\varepsilon(0+,t)]^2 + 2\,[\eta_\varepsilon(0-,t)]^2 \\ \leq 2\,C \int_0^{1/2} \left([\eta_\varepsilon(s,t)]^2 + [\partial_s\eta_\varepsilon(s,t)]^2\right) ds + 2\,C \int_{-1/2}^0 \left([\eta_\varepsilon(s,t)]^2 + [\partial_s\eta_\varepsilon(s,t)]^2\right) ds \,, \end{multline*} where~$C$ is a constant coming from the Sobolev embedding theorem. Recall that~$\varphi_\varepsilon$ is constant on~$(-1,0)$ and $\varphi_\varepsilon \in H^1((0,1/2)) \hookrightarrow C^0([0,1/2])$; more specifically, the first inequality of~\eqref{varphi'.plus} and~\eqref{varphi.plus} respectively yield \begin{equation}\label{varphi.H1} \int_0^{1/2} {\varphi_\varepsilon'}^2 \, ds \leq \frac{C}{m_0 \, \rho_\varepsilon(1/2)} \,, \qquad \int_0^{1/2} \varphi_\varepsilon^2 \, ds \leq \frac{1}{m_0 \, \rho_\varepsilon(1/2)} \,. \end{equation} Integrating with respect to~$t$ above, we deduce $$ [\varphi_\varepsilon(0-) - \varphi_\varepsilon(0+)]^2 \leq 2\,C \int_{\Omega_+'} \left[\eta_\varepsilon^2 + (\partial_s\eta_\varepsilon)^2\right] ds \, dt + 2\,C \int_{\Omega_-'} \left[\eta_\varepsilon^2 + (\partial_s\eta_\varepsilon)^2\right] ds \, dt \,. $$ Applying~\eqref{eta.plus}, the second inequality of~\eqref{varphi'.plus} and~\eqref{eta.minus}, we may write \begin{equation}\label{trace.trick} [\varphi_\varepsilon(0-) - \varphi_\varepsilon(0+)]^2 \leq C \,, \end{equation} where~$C$ is a constant (different from the above) independent of~$\varepsilon$, provided that~\eqref{Ass.small} holds. Finally, applying~\eqref{varphi.H1} and the Sobolev embedding $H^1((0,1/2)) \hookrightarrow C^0([0,1/2])$, we deduce from~\eqref{trace.trick} the following improvement upon~\eqref{varphi.minus} \begin{equation}\label{varphi.minus.improve} \varphi_\varepsilon^2 \leq C \qquad \mbox{on $\Omega_-$} \,. \end{equation} \subsection{As \texorpdfstring{$\varepsilon \to 0$}{epsilonto0} only \texorpdfstring{$\Omega_+$}{Omega+} matters: convergence of eigenvalues and eigenfunctions} Estimate \eqref{varphi.minus.improve}~provides a crucial information whose significance consists in that what happens in~$\Omega_-$ is insignificant. \begin{proposition}\label{Prop.norm} One has $$ \|\psi_{\varepsilon_i}\| \xrightarrow[i \to +\infty]{} \|\varphi_0\|_{L^2_{\gamma_0}((0,+\infty))} \,. $$ \end{proposition} \begin{proof} We have \begin{align*} \|\psi_{\varepsilon}\|^2 = & \int_{\Omega_-} \varphi_{\varepsilon}^2 \, \gamma_\varepsilon \, a_\varepsilon \, \frac{f_\varepsilon \circ g_\varepsilon}{\varepsilon} \, ds \, dt + \int_{\Omega_-} \eta_\varepsilon^2 \, \gamma_\varepsilon \, a_\varepsilon \, \frac{f_\varepsilon \circ g_\varepsilon}{\varepsilon} \, ds \, dt \\ & + \int_{\Omega_-} 2 \, \varphi_{\varepsilon} \, \eta_\varepsilon \, \gamma_\varepsilon \, a_\varepsilon \, \frac{f_\varepsilon \circ g_\varepsilon}{\varepsilon} \, ds \, dt \\ & + \int_{\Omega_+} \varphi_{\varepsilon}^2 \, \gamma_\varepsilon \, ds \, dt + \int_{\Omega_+} \eta_\varepsilon^2 \, \gamma_\varepsilon \, ds \, dt + \int_{\Omega_+} 2 \, \varphi_{\varepsilon} \, \eta_\varepsilon \, \gamma_\varepsilon \, ds \, dt \,. \end{align*} The right hand side of the first line together with the mixed term on the second line goes to zero as $\varepsilon \to 0$. Indeed, recalling~\eqref{convexity}, \eqref{convexity.corol.lower} and $\gamma_\varepsilon \leq 1$, $$ \int_{\Omega_-} \varphi_{\varepsilon}^2 \, \gamma_\varepsilon \, a_\varepsilon \, \frac{f_\varepsilon \circ g_\varepsilon}{\varepsilon} \, ds \, dt \leq 2 \, a_\varepsilon \, \varphi_{\varepsilon}^2 \int_{\Omega_-} (s+1) \, ds \, dt \xrightarrow[\varepsilon \to 0]{} 0 $$ due to~\eqref{varphi.minus.improve}; $$ \int_{\Omega_-} \eta_{\varepsilon}^2 \, \gamma_\varepsilon \, a_\varepsilon \, \frac{f_\varepsilon \circ g_\varepsilon}{\varepsilon} \, ds \, dt \leq 2 \, a_\varepsilon \int_{\Omega_-} \eta_{\varepsilon}^2 \, (s+1) \, ds \xrightarrow[\varepsilon \to 0]{} 0 $$ due to~\eqref{eta.minus}; and the mixed term goes to zero by the Schwarz inequality. Similarly, recalling~\eqref{2-bound.new}, $$ \int_{\Omega_+} \eta_{\varepsilon}^2 \, \gamma_\varepsilon \, ds \, dt \leq c_\varepsilon \int_{\Omega_+} \eta_{\varepsilon}^2 \, \rho_\varepsilon \gamma_0 \, ds \, dt \xrightarrow[\varepsilon \to 0]{} 0 $$ due to~\eqref{eta.plus}; while the Schwarz inequality yields $$ \left| \int_{\Omega_+} 2\, \varphi_{\varepsilon} \, \eta_\varepsilon \, \gamma_\varepsilon \, ds \, dt \right| \leq 2 c_\varepsilon \sqrt{ \int_{\Omega_+} \eta_{\varepsilon}^2 \, \rho_\varepsilon \gamma_0 \, ds \, dt } \sqrt{ \int_{\Omega_+} \varphi_{\varepsilon}^2 \, \rho_\varepsilon \gamma_0 \, ds \, dt } \xrightarrow[\varepsilon \to 0]{} 0 \,, $$ where the second square root is bounded in~$\varepsilon$ due to~\eqref{varphi.plus}. Finally, we write $$ \int_{\Omega_+} \varphi_{\varepsilon_i}^2 \, \gamma_{\varepsilon_i} \, ds \, dt = \int_{\Omega_+} \varphi_{\varepsilon_i}^2 \, \rho_{\varepsilon_i} \gamma_0 \, ds \, dt + \int_{\Omega_+} \varphi_{\varepsilon_i}^2 \, (\gamma_{\varepsilon_i}-\rho_{\varepsilon_i} \gamma_0) \, ds \, dt $$ and observe that the first term on the right hand side tends to the desired result $\|\varphi_0\|_{L^2_{\gamma_0}((0,+\infty))}^2$ as $i \to+ \infty$ by the strong convergence~\eqref{strong.limit}, while the second term vanishes in the limit. In more detail, \begin{multline*} \left| \int_{\Omega_+} \varphi_{\varepsilon_i}^2 \, (\gamma_{\varepsilon_i}-\rho_{\varepsilon_i} \gamma_0) \, ds \, dt \right| = \left| \int_{\Omega_+} \left( \varphi_{\varepsilon_i}^2 \, \rho_{\varepsilon_i} - \varphi_{0}^2 + \varphi_{0}^2 \right) \left(\frac{\gamma_{\varepsilon_i}}{\rho_{\varepsilon_i}}-\gamma_0\right) \, ds \, dt \right| \\ \leq \int_{\Omega_+} \left| \varphi_{\varepsilon_i}^2 \, \rho_{\varepsilon_i} - \varphi_{0}^2 \right| \left(c_{\varepsilon_i}\gamma_0+\gamma_0\right) \, ds \, dt + \int_{\Omega_+} \varphi_{0}^2 \left(\frac{\gamma_{\varepsilon_i}}{\rho_{\varepsilon_i}}-\gamma_0\right) \, ds \, dt \,, \end{multline*} where the the first term after the inequality tends to zero as $i \to +\infty$ by the strong convergence again, while the second term vanishes by the dominated convergence theorem. \end{proof} It follows from Proposition~\ref{Prop.norm} that $\varphi_0 \not= 0$, so that it is indeed an eigenfunction of~$T_0$ due to~\eqref{1Dproblems}. In particular, $\mu_0^+ = \mu_0^-$. Now, let~$\hat{\psi}_\varepsilon$ be a normalised eigenfunction corresponding to possibly another eigenvalue $\hat{\mu}_\varepsilon := \mu_m(\varepsilon)$. Again, we use the decompositions~\eqref{dec.plus} and~\eqref{dec.minus} and distinguish the individual components by tilde. In the same way as we proved Proposition~\ref{Prop.norm}, we can establish \begin{proposition}\label{Prop.inner} One has $$ (\psi_{\varepsilon_i},\hat{\psi}_{\hat\varepsilon_j}) \xrightarrow[i,j \to +\infty]{} (\varphi_0,\hat{\varphi}_0)_{L^2_{\gamma_0}((0,+\infty))} \,. $$ \end{proposition} If $m \not= n$, then $(\psi_{\varepsilon_i},\hat{\psi}_{\hat\varepsilon_j})=0$ and thus $(\varphi_0,\hat{\varphi}_0)_{L^2_{\gamma_0}((0,+\infty))}=0$. Hence~$\varphi_0$ and~$\hat{\varphi}_0$ correspond to distinct eigenvalues of~$T_0$. In particular, $\varphi_0$~is an eigenfunction corresponding to the $(n+1)^\mathrm{th}$ eigenvalue~$\nu_n$ of~$T_0$. Since we get this result for \emph{any} weak limit point of $\{\varphi_\varepsilon\}_{\varepsilon>0}$, we have the convergence results actually in $\varepsilon \to 0$ (no need to pass to subsequences). This completes the proof of Theorem~\ref{Thm.convergence}. \vskip 1cm \noindent \textsc{Acknowledgements}. This paper was partially supported by the grants PRIN 2012 ``Elliptic and parabolic partial differential equations: geometric aspects, related inequalities, and applications'', FIRB 2013 ``Geometrical and qualitative aspects of PDE's'', and STAR 2013 ``Sobolev-Poincar\'e inequalities: embedding constants, stability issues, nonlinear eigenvalues'' (SInECoSINE). {D.K.\ was partially supported by the project RVO61389005 and the GACR grant No.\ 14-06818S.}
\section{Introduction} `Matrix theory' or `matrix model', the theory of $\mathcal{N}=16$ supersymmetric $SU(N)$ gauged matrix quantum mechanics, was proposed in \cite{Banks:1996vh} as a nonperturbative formulation of M-theory. Genuine tests of the BFSS proposal, that is tests which are not guaranteed to work solely by virtue of supersymmetric non-renormalization theorems, have been performed using Monte Carlo methods in a regime where the matrix quantum mechanics is strongly coupled. On the other hand the BFSS proposal can be understood within the framework of gauge/gravity duality: the holographic dual of matrix theory is a lightlike compactification of M-theory in an $SO(9)$-symmetric pp-wave background; moreover compactification to ten dimensions leads to an alternative interpretation whereby weakly-coupled IIA string theory in the near-horizon limit of $N$ $D0$ branes is the holographic dual of $SU(N)$ matrix theory. The gauge/gravity correspondence thus allows one to probe the strong-coupling limit of matrix theory using classical IIA supergravity in a conformal $AdS_2$ times $S^8$ background, which is the near-horizon geometry of D0 branes. This background can be thought of as the uplift to ten dimensions of a domain-wall solution of an effective two-dimensional dilaton-gravity theory. The latter theory is in fact a consistent truncation of IIA supergravity and can thus in principle be used to compute correlation functions in the matrix model involving the operators dual to the graviton and the dilaton, along the lines of holography for non-conformal branes \cite{Itzhaki:1998dd,Boonstra:1998mp,Kanitscheider:2008kd,Kanitscheider:2009as}. However since in two dimensions the dilaton and the graviton can both be gauged away at the classical level, one expects that the corresponding correlation functions should be trivial; we will see that this is indeed consistent with the results of the present paper. To go beyond trivial correlation functions one would need a two-dimensional consistent truncation of IIA which keeps more fields than just the metric and the dilaton. Although an effective lower-dimensional theory is not necessary for holography~\cite{Skenderis:2006uy}, it can help streamline the holographic computations along the lines of holographic renormalisation \cite{Bianchi:2001de,Bianchi:2001kw,Skenderis:2002wp}. Recently a maximally-supersymmetric two-dimensional $SO(9)$ gauged supergravity was constructed in \cite{Ortiz:2012ib}. This theory is expected to be a consistent truncation of IIA supergravity on $S^8$. Subsequently in \cite{Anabalon:2013zka} it was shown that a $U(1)^4$ truncation of the full $SO(9)$ gauge group is indeed a consistent truncation of IIA, and the uplift to ten dimensions was explicitly constructed. In particular the conformal $AdS_2$ times $S^8$ near-horizon geometry was recovered as the uplift to ten dimensions of a supersymmetric domain-wall solution of the two-dimensional theory with sixteen supercharges. In the present paper we will use the half-supersymmetric domain-wall solution of the two-dimensional supergravity to compute correlation functions in the strongly-coupled matrix model using the prescription of holographic renormalization. In particular we compute two-point functions for the operators dual to scalars transforming in the ${\bf 44}$ and the ${\bf 84}$ of $SO(9)$.\footnote{The scalar sector of the two-dimensional maximally supersymmetric $SO(9)$ gauged supergravity contains, besides the dilaton, scalar fields transforming in the ${\bf 44}\oplus{\bf 84}$ of $SO(9)$; its $U(1)^4$ truncation contains the dilaton, four scalars coming from the ${\bf 44}$ and four scalars from the ${\bf 84}$ of $SO(9)$.} Our results are in agreement with the two-point functions previously computed both holographically, from the Kaluza-Klein spectrum of eleven-dimensional supergravity on $S^8$ \cite{Sekino:1999av,Sekino:2000mg}, and directly in the matrix model by Monte Carlo methods \cite{Hanada:2011fq}. Furthermore we construct a half-supersymmetric `deformed' domain-wall solution of two-dimensional $SO(9)$ supergravity which uplifts to an eleven-dimensional pp-wave with symmetry broken from $SO(9)$ to $SO(3)\times SO(6)$. To achieve this deformation we must consider $SO(3)\times SO(6)$-preserving profiles for the scalar fields that go beyond the $U(1)^4$ truncation. As it turns out the resulting eleven-dimensional pp-wave is not of the form of the holographic dual to the BMN matrix model \cite{Berenstein:2002jq} which preserves $\mathcal{N}=32$ supersymmetry;\footnote{It is well-known that all pp-waves of eleven-dimensional supergravity preserve at least sixteen supercharges. The maximally supersymmetric pp-wave \cite{KowalskiGlikman:1984wv} can be thought of as the Penrose limit of either the $AdS_7\times S^4$ or the $AdS_4\times S^7$ background \cite{figuroa}, while there are pp-waves with various possible fractions of supersymmetry between $\mathcal{N}=16$ and $\mathcal{N}=32$ \cite{Gauntlett:2002cs}.} nor does it belong to the class of bubbling M-theory geometries of \cite{Lin:2004nb}. Rather we will show that this $SO(3)\times SO(6)$ deformation should be identified holographically with a vev deformation of the BFSS matrix model. As in the undeformed case we use holographic renormalization to compute two-point correlation functions of operators dual to the scalar fields in the ${\bf 44}$ of $SO(9)$. More precisely, under $SO(3)\times SO(6)$ the ${\bf 44}$ decomposes as ${\bf(1,1)\oplus(1,20)\oplus(5,1)\oplus(3,6)}$; the three distinct two-point functions that we compute in the present paper are those of operators dual to the scalars outside the ${\bf (1,1)}$ singlet.\footnote{This choice was made for simplicity, since the singlet would mix already at the quadratic level with the operators coming from the other representations.} We have checked numerically that in the UV-limit all three reduce to the two-point function of the ${\bf 44}$ scalar computed in the undeformed matrix model. This is consistent with the fact that the deformed domain-wall solution reduces in the limit of small radial direction to the undeformed domain wall. Equivalently it can be checked that the ten-dimensional uplift of the deformed domain-wall solution is asymptotically conformal $AdS_2$ times $S^8$. The plan of the remainder of the paper is as follows. Section~\ref{sec:BFSS} discusses holographic renormalization for the two-dimensional maximal $SO(9)$ supergravity dual to the BFSS matrix quantum mechanics. As a warm-up we compute one- and two-point functions for the operators dual to the graviton and the dilaton and show that they are trivial as expected. We then extend the computation to one- and two-point functions in the scalar sector, where we reproduce the expected field theory results for the corresponding operators. In section \ref{sec:deformation} we construct a half-supersymmetric domain-wall solution of supergravity which breaks $SO(9)$ down to $SO(3)\times SO(6)$ and is expected to provide a holographic description of a corresponding vev deformation of the matrix model. We set up the holographic renormalization around this background and in particular compute the deformed correlation functions in the scalar sector. Some future directions are discussed in section~\ref{sec:discussion}. In appendix \ref{app:validity} we review the various holographic dualities of the matrix model and their respective regimes of validity. In appendix~\ref{app:amb} we review the ambiguity in the holographic dictionary for scalar fields in a certain mass range which will be relevant for our model. \section{BFSS and holographic renormalization} \label{sec:BFSS} In this section, we will employ the effective two-dimensional supergravity that describes fluctuations around the D0-brane near-horizon geometry, and apply the procedure of holographic renormalization in order to extract one- and two-point correlation functions of the corresponding operators in the dual matrix quantum mechanics. \subsection{Effective 2d supergravity and fluctuation equations} The two-dimensional maximally supersymmetric $SO(9)$ supergravity constructed in~\cite{Ortiz:2012ib} describes fluctuations around the $S^8$ compactification of IIA supergravity. The full theory carries a dilaton $\rho$ and 128 scalar fields, transforming as ${\bf 44\oplus 84}$ under $SO(9)$\,. Here, we will only consider its $U(1)^4$ truncation which apart from $\rho$ and the $U(1)^4$ gauge fields carries four more dilaton fields $u_{a}$ from the ${\bf 44}$ and four axion fields $\phi_{a}$ from the ${\bf 84}$ of $SO(9)$\,. The truncated action is given by~\cite{Anabalon:2013zka} \begin{eqnarray} {\cal{L}}&= & -\frac{1}{4} e \rho \, R + \frac{1}{2} e \rho \sum_{a} \partial_{\mu} u_{a} \, \partial^{\mu} u_{a} + \frac{1}{2} e \rho^{1/3} X_{0}^{-1} \sum_{a=1}^{4} X_{a}^{-2} \left( \partial_{\mu} \phi^{a} \right) \left( \partial^{\mu} \phi^{a} \right) \nonumber\\ &&{}- \frac{\rho}{8} \, \varepsilon^{\mu \nu} F_{\mu\nu}^{a} \; y^{a} - e \,V_{\text{pot}}\;, \label{Ltrunc} \end{eqnarray} where we have defined $X_{0} \equiv \prod_a X_a^{-2} $, the scalar kinetic term is defined via \begin{eqnarray} X_{a} \equiv e^{-2\,A_{ab} u_{b}}\;,\qquad A ~\equiv~ {\footnotesize \begin{pmatrix} 1/6 & -1/\sqrt{2} & -1/\sqrt{6} & -1/(2\sqrt{3})\\ 1/6 & 0 & 0 & \sqrt{3}/2\\ 1/6 & 0 & \sqrt{2/3} & -1/(2\sqrt{3})\\ 1/6 & 1/\sqrt{2} & -1/\sqrt{6} & -1/(2\sqrt{3})\\ \end{pmatrix}} \;, \end{eqnarray} and the abelian field strengths $F_{\mu\nu}^{a} \equiv 2 \, \partial^{\vphantom{a}}_{[\mu} A_{\nu]}^{a}$ couple to four auxiliary scalar fields~$y^a$ that can be integrated out from the action. The scalar potential of (\ref{Ltrunc}) is given by \begin{eqnarray} V_{\text{pot}} &=& \rho^{5/9} \,\Big[ \frac{1}{8} \,\Big({X_{0}}^{2} - 8 \sum_{a<b} X_{a} X_{b} - 4 X_{0} \sum_{a} X_{a} \Big) + \frac{1}{2} \, \rho^{-2/3} \sum_{a} X_{a}^{-2} \left( X_{0} - 4 X_{a} \right) (\phi^{a})^{2} \nonumber\\ &&\qquad\quad{} + 2 \, \rho^{-4/3} \sum_{a<b}X_{a}^{-2} X_{b}^{-2} (\phi^{a})^{2} (\phi^{b})^{2} + \frac18 \, \rho^{-2} \sum_{a} X_{a} \, \Big(\rho \, y^{a} + 8 \, \prod_{b \neq a} \phi^{b} \Big)^{2} \nonumber\\ &&\qquad\quad{} + \frac12 \, \rho^{-8/3} X_{0}^{-1} \Big( \sum_{a} \, \rho \,y^{a} \phi^{a} + 8 \prod_{a} \phi^{a} \Big)^{2} \, \Big] \;, \end{eqnarray} as a fourth order polynomial in the scalars $\phi^a$\,. The action (\ref{Ltrunc}) admits a half supersymmetric domain wall solution, in which all scalars and gauge fields vanish and metric and dilaton are given by \begin{eqnarray} \label{dw} ds^2 &=& r^7 dt^2 - dr^2\;,\qquad \rho(r)~=~ r^{9/2}\;. \end{eqnarray} This two-dimensional solution can be uplifted into type IIA supergravity as \begin{eqnarray} ds_{10}^2 = r^{-7/8} \big( r^7 \text{dt}^2 - (\text{dr}^2 + r^2 \, d\Omega_{8}^2)\big) \;,\quad \Phi = -\frac{21}{8} \ln r\;,\quad F = d\, \big( r^7 \text{dt} \big) \;, \label{abh} \end{eqnarray} (with 10D dilaton $\Phi$ and two-form flux $F$) and further to an eleven-dimensional pp-wave solution \cite{Hull:1984vh,Townsend:1998qp,Gauntlett:2002cs} \begin{equation}\label{2.6} ds_{11}^{2}= dx^+ \, dx^- + (1-r^{-7}) (dx^-)^2 - (dr^2+ r^2 \,d\Omega_{8}^{2}) \,. \end{equation} In this section, we will compute correlation functions associated to the quadratic fluctuations around the domain wall (\ref{dw}). Since scalars originating from different $SO(9)$ representations do not mix at the quadratic level, we will only need the truncated action (\ref{Ltrunc}) of two-dimensional dilaton gravity coupled to one of the scalars $X_a$ and one of the scalars $\phi^a$. We will denote these two scalars by $y_{44}$ and $y_{84}$ respectively (referring to their $SO(9)$ origin), and collectively by $y_n$. Moreover, it will be convenient to go to a frame in which the background metric of (\ref{dw}) becomes pure AdS which is achieved by rescaling the fields as \begin{equation} t\rightarrow \frac{2}{5} \, t \,,\quad r \rightarrow \,r^{-1/5}\,, \quad g_{\mu \nu} \rightarrow \frac{4}{25} \,\rho^{4/9} \, g_{\mu \nu}\,. \end{equation} In this frame, and after Wick rotation to Euclidean signature, the action takes the canonical form~\cite{Kanitscheider:2008kd} \begin{equation} \label{effy} S = \frac{1}{4} \, \int d^{2}x\, \sqrt{|g|} \, e^{\gamma \phi} \left( R + \beta \left(\partial \phi \right)^{2} + C - e^{a_n\phi} \left( \left( \partial y_n \right)^{2} - m_n^2 \, y_n^{2} \right) \right) . \end{equation} with $\rho\equiv e^{\gamma \phi}$, and the constants \begin{equation} \gamma \equiv -\frac{6}{7} \,,\quad \beta\equiv\frac{16}{49} \,,\quad C\equiv\frac{126}{25}\,, \end{equation} describing the dilaton-gravity sector. With these coordinates, the boundary of AdS is located at ${r=0}$ and the background (\ref{dw}) takes the form \begin{equation} \label{backG} ds^2 = \frac{1}{r} dt^2 + \frac{1}{4r^2}dr^2\,,\qquad e^\phi = r^{\alpha} \,,\qquad \alpha\equiv \frac{21}{20}\;. \end{equation} The scalar couplings in (\ref{effy}) are characterized by the constants $a_n$ and $m_n$ which take different values for the scalars in the 44 and 84, respectively: \begin{eqnarray} \label{amy} &&{} a_{44}\equiv0 \,, \quad m_{44}^2\equiv\frac{8}{5}\,,\;\; \quad y_{44} \equiv 6\sqrt{2} \,x \,, \quad\mbox{with}\quad X_{1,2,3,4}=e^{-2x}\;, \nonumber\\ &&{} a_{84}\equiv \frac{4}{7} \,, \quad m_{84}^2\equiv\frac{12}{25}\,, \quad y_{84} \equiv \sqrt{2} \,\phi^{a=1} \,. \end{eqnarray} Let us note that the addition of scalar matter in (2.8) is the source of some technical complications with respect to the standard treatment of the dilaton gravity sector~\cite{Kanitscheider:2008kd,Kanitscheider:2009as}. In particular the fact that the scalars $y_{84}$ arise with a non-vanishing relative dilaton power $a_{84}$ prevents us from using the methods of~\cite{Kanitscheider:2009as} and translate the non-conformal holographic problem into a pure AdS background in some suitable higher dimension. However, it is straightforward to extend the analysis of \cite{Kanitscheider:2008kd} to the presence of additional matter fields. The equations of motion follow from (\ref{effy}) and yield \begin{align} \label{eomy} 0&= \left( \nabla_{\mu} \partial_{\nu} \phi \right) -\frac{g_{\mu \nu}}{2} \nabla \partial \phi - \Big( \frac{\beta}{\gamma} - \gamma \Big) \Big( \left( \partial_{\mu} \phi \right) \left( \partial_{\nu} \phi \right) - \frac{g_{\mu \nu}}{2} \left( \partial \phi \right)^2 \Big) \nonumber\\ &\quad+\frac{e^{a_n\phi}}{\gamma} \big( \partial_{\mu} y_n \partial_{\nu} y_n - \frac{1}{2} g_{\mu \nu} (\partial y_n)^2 \big)\,,\nonumber\\ 0&= \gamma \nabla \partial \phi + \gamma^2 \big( \partial \phi \big)^2 - C - m_n^2 e^{a_n\phi} y_n^2\,,\nonumber\\ 0&= R - 2\frac{\beta}{\gamma} \nabla \partial \phi - \beta \left( \partial \phi \right)^{2} + C - \big(1+\frac{a_n}{\gamma} \big) e^{a_n\phi} \left( \left( \partial y \right)^{2} - m_n^2 \, y_n^{2} \right) \,,\nonumber\\ 0&= \nabla^{\mu} \big( e^{(a_n+\gamma)\phi}\, \partial_{\mu} y_n \big) + m_n^2 e^{(a_n+\gamma)\phi}\, y_n\,. \end{align} They respectively stand for: the traceless and trace part of Einstein equations, the dilaton field equation, and the scalar equations of motion. \subsection{Asymptotic expansions} Following the procedure of holographic renormalization \cite{Bianchi:2001de, Bianchi:2001kw,Skenderis:2002wp,Kanitscheider:2008kd}, we first compute the asymptotic expansions of all fields at the boundary $r=0$. As an illustration, let us first restrict to the dilaton-gravity sector, i.e.\ set all scalar fields other than the dilaton to zero, in which case we reproduce the results of \cite{Kanitscheider:2008kd} for the (degenerate) case of the D0 branes. The fluctuation ansatz for metric and dilaton is given by \begin{equation} \begin{aligned} ds^2 &= \frac{f(t,r)}{r} dt^2 + \frac{1}{4r^2}dr^2\,,\\ \phi &= \alpha\, \ln r + \frac{\kappa(t,r)}{\gamma}\,.\\ \end{aligned} \end{equation} with functions $f(t,r)$, $\kappa(t,r)$ admitting a (fractional) power expansion in $r$ near $r=0$ \begin{eqnarray} f(t,r)= f_{(0)}(t) + \underset{r \to 0}{o}(1) \,,\qquad \kappa(t,r)= \kappa_{(0)}(t) + \underset{r \to 0}{o}(1) \,. \end{eqnarray} According to the equations of motion \eqref{eomy}, the functions $f(t,r)$ and $\kappa(t,r)$ are subject to the non-linear partial differential equations \begin{align} \label{efk} 0&= - \frac{1}{4} \big( f^{-1} f' \big)^{2} + \frac{1}{2} f^{-1} f'' + \kappa'' + \big( 1-\frac{\beta}{\gamma^{2}}\big) \big( \kappa'\big)^{2} \,, \nonumber\\ 0&=\big(1-\frac{\beta}{\gamma^{2}} \big) \dot{\kappa} \kappa' + \dot{\kappa}' - \frac{1}{2} f' f^{-1} \dot{\kappa} \,,\\ 0&=2 \alpha \gamma f' + r \big( 2 f'' - f^{-1} \big( f' \big)^{2}\big) + \ddot{\kappa} - \frac{1}{2} f^{-1} \dot{f} \dot{\kappa} + \big(1-\frac{\beta}{\gamma^{2}} \big)\big( \dot{\kappa}\big)^{2} - 2f \big( 1 - r f^{-1} f'\big) \kappa' \,,\nonumber\\ 0&=4r \big( \kappa'' + \big( \kappa' \big)^{2} \big) + \big( 8 \alpha \gamma + 2 + 2rf^{-1} f' \big) \kappa'+ f^{-1} \big( \ddot{\kappa} - \frac{1}{2} f^{-1} \dot{f} \dot{\kappa}+ \big( \dot{\kappa}\big)^{2} \big) + 2 f^{-1} f' \alpha \gamma \,,\nonumber \end{align} where dots and primes refer to $\partial_t $ and $\partial_r$, respectively. Closer inspection of these equations shows that its solutions admit a fractional power expansion around $r=0$ % \begin{align} f(t,r)&=f_{(0)}(t) + r \, f_{(5)}(t)+ r^{\sigma} \, f_{(5\sigma)}(t) + \dots \,,\nonumber\\ \kappa(t,r)&=\kappa_{(0)}(t) + r \, \kappa_{(5)}(t)+ r^{\sigma} \, \kappa_{(5\sigma)}(t) + \dots \,, \label{fk0} \end{align} where $\sigma = \frac{1}{2} - \alpha \gamma=\frac{7}{5}$ denotes the first non-integer power in the expansion, whose coefficient is not determined by the equations of motion (\ref{efk}). In generic dimensions, this coefficient carries the information about the two-point correlation functions of the associated operators. In two dimensions (i.e.\ for the $p=0$ branes) this structure is highly degenerate. Specifically, the equations of motion (\ref{efk}) determine the coefficients $\kappa_{(5)}$, $f_{(5)}$ as \begin{align} \label{f5} \kappa_{(5)}&=\frac{5}{36} \, f_{(0)}^{-1} {\dot{\kappa}}_{(0)}^{2} \,,\nonumber\\ f_{(5)} &= \frac{5}{9} \big( {\ddot{\kappa}}_{(0)} - \frac{1}{2} f_{(0)}^{-1} {\dot{f}}_{(0)} {\dot{\kappa}}_{(0)} + \frac{5}{18} {\dot{\kappa}}_{(0)}^{2} \big) \,, \end{align} and constrain the coefficients $\kappa_{(5\sigma)}$, $f_{(5\sigma)}$ as \begin{align} \label{fs} 0&=f_{(5\sigma)} + 2 f_{(0)} \kappa_{(5\sigma)} \,,\nonumber\\ 0&=\dot{\kappa}_{(5\sigma)} + \frac{14}{9} \dot{\kappa}_{(0)} \kappa_{(5\sigma)} \,. \end{align} The latter conditions imply the two-dimensional analogue of what in higher dimensions expresses the diffeomorphism and trace Ward identities \cite{Bianchi:2001kw,Kanitscheider:2008kd}. In two dimensions these contraints imply that there are no non-trivial correlation functions associated to the operators dual to $f$ and $\kappa$, respectively, as we shall discuss shortly. This is related to the fact that in two dimensions the dilaton-gravity sector does not carry any propagating degrees of freedom. In this case, the interesting structure is sitting in the scalar sector of the theory. Let us thus repeat the previous analysis in presence of the scalar fields. Consider first the action (\ref{effy}) with scalar fields from the ${\bf 44}$ and the ${\bf 84}$ of $SO(9)$. The equations of motion obtained from variation of (\ref{effy}) then imply a generalization of the ansatz (\ref{fk0}) to a fractional expansion of the type \begin{align} \label{ax} f(t,r)&=f_{(0)}(t) + r^{4/5} \, f_{(4)}(t)+ r \, f_{(5)}(t) + r^{7/5} \,f_{(7)}(t) + \dots\;,\nonumber\\ \kappa(t,r)&=\kappa_{(0)}(t) + r^{4/5} \, \kappa_{(4)}(t) + r \, \kappa_{(5)}(t) + r^{7/5} \, \kappa_{(7)}(t) + \dots\;,\nonumber\\ y_{44}(r,t)&= r^{2/5} \, x_{(2)}(t) + r \, x_{(5)}(t) + \dots \;,\nonumber\\ y_{84}(r,t)&= r^{1/5} \, y_{(1)}(t) + r^{3/5} \, y_{(3)}(t) + \dots \;, \end{align} where $x_{(5)}$ and $y_{(3)}$ correspond to the coefficients in the scalar expansion that are left undetermined by the equations of motion. The intermediate coefficients in the series expansion are determined by the equations of motion to \begin{align} \label{cx} \kappa_{(4)} &= - \frac{1}{4} \, x_{(2)}^{2} \,,\nonumber\\ \kappa_{(5)} &=\frac{5}{36} \, f_{(0)}^{-1} {\dot{\kappa}}_{(0)}^{2} - \frac{1}{10} e^{-\frac{2 \kappa_{(0)}}{3}} y_{(1)}^{2} \,,\nonumber\\ \dot{\kappa}_{(7)} &= - \frac{14}{9} \dot{\kappa}_{(0)} \kappa_{(7)} -\frac{e^{-\frac{2}{3}\kappa_{(0)}}}{7} \big(3 \dot{y}_{(1)} y_{(3)} + y_{(1)} \dot{y}_{(3)} + \frac{4}{3} y_{(1)} y_{(3)} \dot{\kappa}_{(0)} \big) \nonumber\\ &\quad - \frac{1}{7} \big( 5 \dot{x}_{(2)} x_{(5)} + 2 x_{(2)} \dot{x}_{(5)} + \frac{40}{9} x_{(2)} x_{(5)} \dot{\kappa}_{(0)} \big) \,,\nonumber\\ f_{(4)} &= - \frac{5}{18} f_{(0)} \, x_{(2)}^{2}\,,\nonumber\\ f_{(5)} &= \frac{5}{9} \big( {\ddot{\kappa}}_{(0)} - \frac{1}{2} f_{(0)}^{-1} {\dot{f}}_{(0)} {\dot{\kappa}}_{(0)} + \frac{5}{18} {\dot{\kappa}}_{(0)}^{2} \big) + \frac{1}{45} e^{-\frac{2 \kappa_{(0)}}{3}} f_{(0)} y_{(1)}^{2}\,, \nonumber\\ f_{(7)} &= -2 f_{(0)} \kappa_{(7)} - \frac{80}{63} f_{(0)} \, x_{(2)} x_{(5)} - \frac{8}{21} e^{-\frac{2 \kappa_{(0)}}{3}} f_{(0)} y_{(1)} y_{(3)}\,. \end{align} In absence of the scalar fields these expressions consistently reproduce (\ref{f5}). \subsection{Regularization and counterterms} \paragraph{On-shell action} The central object for the computation of correlation functions is the action (\ref{effy}) evaluated on-shell. Using the dilaton field equation from (\ref{eomy}), the on-shell Lagrangian reduces to \begin{eqnarray} {\cal L}|_{\rm on-shell} &=& \frac{2\beta}{\gamma} \sqrt{|\mathrm{det}g|} \,\nabla \big( e^{\gamma \phi} \partial \phi \big) +\frac{a_{84}}{\gamma} \sqrt{|\mathrm{det}g|} \,e^{a_{84} \phi} \left((\partial y_{84})^2-m^2 y_{84}^2\right)\,. \label{Lonshell1} \end{eqnarray} Note that no explicit scalar dependence on $y_{44}$ appears in the Lagrangian. This is due to the fact that these scalars appear coupled with the same dilaton power as the Einstein-Hilbert term, c.f.~(\ref{effy}), (\ref{amy}), thus disappear form the action upon using the dilaton equation of motion. Moreover, we need to add the Gibbons-Hawking term in order to take into account the boundary of the background spacetime \begin{equation} \int_{{\cal M}} d^{2}x \, \sqrt{|\mathrm{det}g|} \,e^{\gamma \phi}\, R \quad \longrightarrow \quad \int_{{\cal M}} d^{2}x \, \sqrt{|\mathrm{det}g|} \,e^{\gamma \phi}\, R + \int_{\partial {\cal{M}}} ds \sqrt{h} \,e^{\gamma \phi}\, 2\, K \,. \end{equation} Here $h$ is the induced metric on the (one-dimensional) boundary and $K$ is the trace of the extrinsic curvature of the boundary that can be computed from a unit length vector $n^{\mu}$ normal to the boundary \begin{equation} K = \nabla_{\mu} n^{\mu}\,. \end{equation} Putting everything together, the full on shell action is given by \begin{equation} \label{AdSEdW0} S_{\text{on-shell}} = \frac{1}{2} \int_{\partial {\cal M}} dt \sqrt{h} \, e^{\gamma \phi} \left( K + \frac{\beta}{\gamma} n^{\mu} \partial_{\mu} \phi + \frac{2}{7\gamma} e^{\frac{4}{7} \phi} \, y_{84} \; n^{\mu} \partial_{\mu} y_{84} \right)\,, \end{equation} where the boundary is located at $r=0$. Because the integral diverges when $r \to 0$, the first step of holographic renormalization consists in regularizing the integral by introducing a parameter $\epsilon$ in order to control the divergences \begin{equation} \label{AdSEdW} S_{\text{reg}} = \frac{1}{2} \int_{\partial {AAdS}, r=\epsilon} dt \sqrt{h} \, e^{\gamma \phi} \left( K + \frac{\beta}{\gamma} n^{\mu} \partial_{\mu} \phi + \frac{2}{7\gamma} e^{\frac{4}{7} \phi} \, y_{84} \; n^{\mu} \partial_{\mu} y_{84} \right)\,. \end{equation} Knowing the asymptotic behaviour of the fields near the boundary, the regularized on-shell action \eqref{AdSEdW} may be evaluated as a function of $\epsilon$. Let us recall that $n^{\mu}$ is a unit vector ($n^{\mu} n_{\mu} = 1$) normal to the boundary \begin{equation} n^{\mu} \partial_{\mu}= n \, \partial_r = 2r\partial_r \,, \end{equation} and \begin{eqnarray} h = \frac{f(t,r)}{r} dt^{2}\,,\qquad K = \nabla_{\mu} n^{\mu} = -1 + r \, \partial_{r} \ln f \,. \end{eqnarray} Inserting the expansion \eqref{ax} in the action \eqref{AdSEdW} leads to the different contributions \begin{align} \sqrt{h} \, e^{\gamma \phi} &= |f_{(0)}|^{1/2} e^{\kappa_{(0)}} \, \epsilon^{-7/5}\, \Big[ 1 + \big(\frac{1}{2} f_{(0)}^{-1} f_{(4)} + \kappa_{(4)} \big) \, \epsilon^{4/5} + \big(\frac{1}{2} f_{(0)}^{-1} f_{(5)} + \kappa_{(5)} \big) \, \epsilon \nonumber\\ &\hspace*{4cm}+ \big( \frac{1}{2} f_{(0)}^{-1} f_{(7)} + \kappa_{(7)} \big) \, \epsilon^{7/5} \Big]+\dots \,,\nonumber\\ \left.K\right|_{r=\epsilon}&= -1 + f_{(0)}^{-1} \, \Big[ \frac{4}{5} f_{(4)} \, \epsilon^{4/5} + f_{(5)} \, \epsilon + \frac{7}{5} f_{(7)} \, \epsilon^{7/5} \Big] +\dots\,,\nonumber\\ \left.n^{\mu} \partial_{\mu} \phi\right|_{r=\epsilon} &= 2\alpha + \frac{2}{\gamma} \Big[ \frac{4}{5} \kappa_{(4)} \, \epsilon^{4/5} + \kappa_{(5)} \epsilon + \frac{7}{5} \kappa_{(7)} \epsilon^{7/5} \Big] +\dots \,,\nonumber\\ \left.e^{\frac{4}{7} \phi} \, y \; n^{\mu} \partial_{\mu} y\right|_{r=\epsilon} &= e^{-\frac{2}{3} \kappa_{(0)}} \, \Big[ \frac{2}{5} \, y_{(1)}^2 \, \epsilon + \frac{4}{5} \, y_{(1)} y_{(3)} \, \epsilon^{7/5} \Big] + \dots \,. \end{align} The most divergent term in this expansion comes from the determinant of the induced metric times the dilaton and involves a global factor of $\epsilon^{-7/5}$. The on-shell action can now be expressed as a perturbative expansion in $r=\epsilon$ up to terms vanishing when $\epsilon$ goes to zero \begin{eqnarray} \label{Sdiv} S_{\text{reg}} &=& \frac{1}{2} \int dt \; |f_{(0)}|^{1/2} e^{\kappa_{(0)}} \left( L_{(-7)}\,\epsilon^{-7/5}+ L_{(-3)}\,\epsilon^{-3/5}+L_{(-2)}\,\epsilon^{-2/5}+L_{(0)}\,\epsilon^{0} + o(1) \right) \;, \nonumber\\[2ex] &&{} L_{(-7)} ~\equiv~ -1 + \frac{2\alpha \beta}{\gamma}~=~ -\frac95 \;,\\ &&{} L_{(-3)} ~\equiv~ -\frac95 \left( \frac{1}{2} f_{(0)}^{-1} f_{(4)} + \kappa_{(4)} \right) + \frac{4}{5} f_{(0)}^{-1} f_{(4)} + \frac{4}{5} \frac{2\beta}{\gamma^{2}} \kappa_{(4)} \;,\nonumber\\ &&{} L_{(-2)} ~\equiv~ -\frac95 \left( \frac{1}{2} f_{(0)}^{-1} f_{(5)} + \kappa_{(5)} \right) + f_{(0)}^{-1} f_{(5)} + \frac{2\beta}{\gamma^{2}} \kappa_{(5)} + \frac{4}{35 \gamma} e^{-\frac{2}{3} \kappa_{(0)}} y_{(1)}^2 \;,\nonumber \\ &&{} L_{(0)} ~\equiv~ -\frac95 \left( \frac{1}{2} f_{(0)}^{-1} f_{(7)} + \kappa_{(7)} \right) + \frac{7}{5} f_{(0)}^{-1} f_{(7)} + \frac{7}{5}\frac{2\beta}{\gamma^{2}} \, \kappa_{(7)} + \frac{16}{35 \gamma} e^{-\frac{2}{3} \kappa_{(0)}} y_{(1)} y_{(3)}\;. \nonumber \end{eqnarray} We note that there is no explicit dependence on the scalars $x_{(2)}$, $x_{(5)}$, c.f.\ the discussion after (\ref{Lonshell1}). The dependence of the regularized action on these fields enters implicitly via the metric and dilaton components (\ref{cx}). \paragraph{Counterterms} The first counter-term required for cancelling the most divergent contribution in (\ref{Sdiv}) takes the form of an exponential dilaton potential \begin{equation} S_{\text{ct1}} = \frac{1}{2} \int \text{dt} \sqrt{h} \, e^{\gamma \phi} \Big( 1 - \frac{2\alpha \beta}{\gamma} \Big)\,. \label{sct1} \end{equation} This kills the first divergent term in \eqref{Sdiv} and also modifies the sub-leading terms \begin{equation} \label{div1} \begin{aligned} S_{\text{reg}} + S_{\text{ct1}} &= \frac{1}{2} \int dt \; |f_{(0)}|^{1/2} e^{\kappa_{(0)}} \Big[ \frac{4}{5} \big( f_{(0)}^{-1} f_{(4)} + \frac{2\beta}{\gamma^{2}} \kappa_{(4)} \big) \;\epsilon^{-3/5} \\ &\quad+\big( f_{(0)}^{-1} f_{(5)} + \frac{2\beta}{\gamma^{2}} \kappa_{(5)} + \frac{4}{35 \gamma} e^{-\frac{2}{3} \kappa_{(0)}} y_{(1)}^2 \big) \;\epsilon^{-2/5} \\ &\quad+ \frac{7}{5} \big( f_{(0)}^{-1} f_{(7)} + \frac{2\beta}{\gamma^{2}} \kappa_{(7)} \big) +\frac{16}{35 \gamma} e^{-\frac{2}{3} \kappa_{(0)}} y_{(1)} y_{(3)} \; +\; o(1)\Big]\,.\\ \end{aligned} \end{equation} Moreover, $f_{(5)}$ and $\kappa_{(5)}$ are related to the sources by \eqref{cx}. This corresponds to the expansion of \begin{equation} \begin{aligned} \left.\big( \nabla^{t} \partial_{t} \phi \big)\right|_{r=\epsilon} &= \frac{f_{(0)}^{-1}}{\gamma} \big( {\ddot{\kappa}}_{(0)} - \frac{1}{2} f_{(0)}^{-1} {\dot{f}}_{(0)} {\dot{\kappa}}_{(0)} \big) \, \epsilon + o(\epsilon)\,,\\ \left.\big( \partial \phi \big)^{2}\right|_{r=\epsilon} &= \frac{f_{(0)}^{-1} {\dot{\kappa}}_{(0)}^{2}}{\gamma^{2}} \, \epsilon + o(\epsilon)\,,\\ \end{aligned} \end{equation} and determines the form of the second counter-term \begin{equation} \begin{aligned} S_{\text{ct2}} &= \frac{1}{2} \int dt \sqrt{h} \, e^{\gamma \phi} \Big( \frac{10}{21} \big( \nabla^{t} \partial_{t} \phi \big) - \frac{10}{49} \big( \partial \phi \big)^{2} \Big)\\ &= \frac{1}{2} \int dt |f_{(0)}|^{1/2} e^{\kappa_{(0)}} \big(-\frac{5}{9} f_{(0)}^{-1}\big)\Big( {\ddot{\kappa}}_{(0)} - \frac{1}{2} f_{(0)}^{-1} {\dot{f}}_{(0)} {\dot{\kappa}}_{(0)} + \frac{1}{2} {\dot{\kappa}}_{(0)}^{2} \Big) \, \epsilon^{-2/5} + o(1)\;,\\ \end{aligned} \end{equation} These terms cancel the $f_{(5)}$ and $\kappa_{(5)}$ contributions to the divergent part of the on-shell action (\ref{div1}). Upon furthermore replacing $f_{(4)}$ and $\kappa_{(4)}$ by their expression from \eqref{cx}, the resulting action reads \begin{equation} \begin{aligned} S_{\text{reg}} + S_{\text{ct1}} + S_{\text{ct2}} &= \frac{1}{2} \int dt \; |f_{(0)}|^{1/2} e^{\kappa_{(0)}} \Big[ -\frac{2}{5} x_{(2)}^2 \;\epsilon^{-3/5} - \frac{1}{5} e^{-\frac{2}{3} \kappa_{(0)}} y_{(1)}^2 \;\epsilon^{-2/5} \\ &\qquad\quad+ \frac{7}{5} \big( f_{(0)}^{-1} f_{(7)} + \frac{2\beta}{\gamma^{2}} \kappa_{(7)} \big) +\frac{16}{35 \gamma} e^{-\frac{2}{3} \kappa_{(0)}} y_{(1)} y_{(3)} \; +\; o(1)\Big]\,.\\ \end{aligned} \end{equation} From this expression we read off the last counterterms for the matter couplings \begin{eqnarray} S_{\text{ct3}} &=& \frac{1}{5} \int dt \sqrt{h} \, e^{\gamma \phi}\, y_{(44)}^{2} \,,\nonumber\\ S_{\text{ct4}} &=& \frac{1}{10} \int dt \sqrt{h} \, e^{(\gamma + a) \phi} \, y_{(84)}^{2} \,. \end{eqnarray} After renormalization by all counter-terms, the on-shell action is given by \begin{eqnarray} S_{\text{ren}} &=& S_{\text{reg}} + S_{\text{ct1}} + S_{\text{ct2}}+ S_{\text{ct3}} + S_{\text{ct4}} \\ &=& \frac{1}{2} \int dt |f_{(0)}|^{1/2} e^{\kappa_{(0)}} \bigg[ \frac{7}{5} \big( f_{(0)}^{-1} f_{(7)} + \frac{2\beta}{\gamma^{2}} \kappa_{(7)} \big) +\frac{4}{5} x_{(2)} x_{(5)} - \frac{2}{15} e^{-\frac{2 \kappa_{(0)}}{3}} y_{(1)} y_{(3)} \bigg] \;. \nonumber \end{eqnarray} and contains only finite terms in the limit $\epsilon\rightarrow0$. Eventually, taking into account the relation between $f_{(7)}$ and $\kappa_{(7)}$ from \eqref{fs}, the renormalized action takes the final form \begin{eqnarray} \label{Sren} S_{\text{ren}} & =& \int \text{dt}|f_{(0)}|^{1/2} e^{\kappa_{(0)}} \left( - \frac{7}{9} \kappa_{(7)} - \frac{22}{45} x_{(2)} x_{(5)} - \frac{1}{3} e^{-\frac{2 \kappa_{(0)}}{3}} y_{(1)} y_{(3)} \right) \;. \end{eqnarray} \subsection{Correlation functions} \label{subsec:correlation} \paragraph{One-point functions} From the renormalized action (\ref{Sren}) we may now extract the one-point correlation functions for the various dual operators by functional derivation. For the operators dual to the dilaton and the two-dimensional metric, we thus obtain \begin{equation} \begin{aligned} \langle {\cal{O}}_{\kappa} (t) \rangle &= \frac{1}{{|f_{(0)}(t)|}^{1/2}} \, \frac{\delta S_{\text{ren}}}{\delta \kappa_{(0)}(t)} = \,e^{\kappa_{(0)}} \,\left( - \frac{7}{9} \kappa_{(7)} - \frac{22}{45} x_{(2)} x_{(5)} - \frac{1}{9} e^{-\frac{2 \kappa_{(0)}}{3}} y_{(1)} y_{(3)} \right) \,,\\ \langle {\cal{O}}_{f} (t) \rangle &= \frac{2}{{|f_{(0)}(t)|}^{1/2}} \, \frac{\delta S_{\text{ren}}}{\delta f_{(0)}^{-1}(t)} =e^{\kappa_{(0)}} \, \left( \frac{7}{9} \kappa_{(7)} + \frac{22}{45} x_{(2)} x_{(5)} + \frac{1}{3} e^{-\frac{2 \kappa_{(0)}}{3}} y_{(1)} y_{(3)} \right) \,.\\ \end{aligned} \end{equation} Similarly, in the matter sector, we derive the following one-point correlation functions for the operators dual to the scalars in the {\bf 44} and the {\bf 84} representation \begin{align} \langle {\cal{O}}_{44} (t) \rangle &= \frac{1}{{|f_{(0)}(t)|}^{1/2}} \, \frac{\delta S_{\text{ren}}}{\delta x_{(2)}(t)} \propto e^{\kappa_{(0)}} \, x_{(5)}(t)\,,\\ \langle {\cal{O}}_{84} (t) \rangle &= \frac{1}{{|f_{(0)}(t)|}^{1/2}} \, \frac{\delta S_{\text{ren}}}{\delta y_{(1)}(t)} \propto e^{\kappa_{(0)}/3} \, y_{(3)}\,. \end{align} \paragraph{Two-point function} The two-point correlation functions are obtained by further functional derivative of the one-point functions. To this end, we first need to determine the dependence of the `response' functions $\{f_{(7)}, \kappa_{(7)}, x_{(5)}, y_{(3)}\}$ on the `source functions' $\{f_{(0)}, \kappa_{(0)}, x_{(2)}, y_{(1)}\}$. This dependence is fixed by the requirement that the solution of the field equations remains regular in the bulk. In absence of an exact solution of the non-linear equations of motion, the two-point correlation functions can be computed from exact solutions of the linearized equations of motion. In the dilaton-gravity sector, linearizing the field equations around the background \begin{align} f(t,r) &= 1+ \eta(t,r) \,,\nonumber\\ \kappa(t,r) &= 0 + \kappa(t,r)\,, \end{align} leads to the set of equations \begin{align} 0&=\frac{1}{2} \eta^{''} + \kappa^{''} \,,\qquad 0=\dot{\kappa}^{'} \,,\nonumber\\ 0&=2 \alpha \gamma \, \eta^{'} + 2 r \, \eta^{''} + \ddot{\kappa} - 2 \kappa^{'} \,,\nonumber\\ 0&=4r \kappa^{''} + \left( 2+ 8 \alpha \gamma\right) \kappa^{'} + \ddot{\kappa} + 2 \alpha \gamma \, \eta^{'}\,, \end{align} whose general solution is provided by \begin{align} \label{linG} \eta(t,r) &= \eta_{(0)}(t) +\frac59\,\ddot{\kappa}_{(0)}(t) \, r - 2 A \, r^{7/5} \,,\nonumber\\ \kappa(t,r)&= \kappa_{(0)}(t) + A \, r^{7/5} \,, \end{align} with real constant $A$. Regularity in the bulk requires that $A=0$ which translates into $f_{(7)}=0=\kappa_{(7)}$\,. As a result, all related two-point correlation functions vanish. \begin{equation} \langle {\cal O}_{\kappa}(t_1) {\cal O}_{\kappa}(t_2) \rangle = 0 = \langle {\cal O}_{f}(t_1) {\cal O}_{f}(t_2) \rangle \;. \end{equation} As alluded to above, this is a consequence of the fact that in two dimensions the dilaton-gravity sector does not carry propagating degrees of freedom. The interesting structure of correlation functions is situated in the matter sector. Linearizing the scalar field equations (\ref{eomy}) around the background (\ref{dw11}) yields a linear differential equation that can be simplified by taking the Fourier transform with respect to time: \begin{equation} \label{scalar} r^{2} \, \tilde{y}_n''(q,r) + \Big(\frac{21}{20} \, a_n - \frac{2}{5} \Big) r \, \tilde{y}_n'(q,r) - \frac{1}{4} ( q^{2} r - m_n^2 ) \, \tilde{y}_n(q,r)=0 \;. \end{equation} For the scalars from the {\bf 44} and the {\bf 84} with the parameters given by (\ref{amy}), the asymptotic analysis of this equation yields an expansion \begin{eqnarray} \tilde{y}_{(44)}(r,q) &=& r^{2/5} \left(\tilde{x}_{(2)}(q) + r^{3/5} \, \tilde{x}_{(5)}(q) +\dots\right) \;,\nonumber\\ \tilde{y}_{(84)}(r,q) &=& r^{1/5} \left(\tilde{y}_{(1)}(q) + r^{2/5} \, \tilde{y}_{(3)}(q) +\dots\right)\;, \label{axy} \end{eqnarray} in accordance with (\ref{ax}). Let us first consider the scalar fields in the {\bf 44}. The corresponding equation (\ref{scalar}) can be brought in a more canonical form by making the following change of variables and redefinitions \begin{equation} \label{can} \tilde{r} = q\, \sqrt{r}\,, \qquad \tilde{y}_{(44)}(q,\tilde{r}) = \tilde{r}^{\lambda} \, s(q,\tilde{r}) \,, \qquad \lambda = \frac{7}{5}\,, \end{equation} upon which the equation becomes \begin{equation} \label{scan} \tilde{r}^{2} s^{''} + \tilde{r} \, s^{'} -\big(\tilde{r}^2 + \lambda^2 -m^2\big) \, s =0\,. \end{equation} This corresponds to the modified Bessel's equation with parameter $\sqrt{\lambda^2-m^2}=\frac35$. It admits two linearly independent solutions which may be described by modified Bessel function of the first kind $I$ and the second kind $K$. The solution regular in the bulk is given by % \begin{eqnarray} \tilde{y}_{(44)}(q,r) &=& \tilde{r}^{7/5} \, \mathrm{Bessel}_K (3/5,\tilde{r})\,, \label{Bess} \end{eqnarray} and we can infer its asymptotic development near $r=0$ as \begin{align} \tilde{y}_{(44)}(q,r) &=q^{4/5} \, \left( \frac{\Gamma (\frac{3}{5})}{2^{2/5}} \, r^{2/5} + \frac{ \Gamma(-\frac{3}{5})}{ 2^{8/5}} \, q^{6/5} \, r + \frac{5 \Gamma(\frac{3}{5})}{ 2^{17/5}} \, q^2 \, r^{7/5} + \underset{r \to 0}{o}(r^{7/5}) \right)\;. \end{align} Comparing to the general expansion (\ref{axy}) we find that \begin{equation}\label{251} {\tilde{x}}_{5}(q) \propto q^{6/5}\, {\tilde{x}}_{2}(q) \,. \end{equation} Before proceeding with the computation of the two-point function, we should recall the possible ambiguity in the assignment of conformal dimensions for the scalar fields discussed in appendix~\ref{app:amb}. The scalar fields in the {\bf 44} precisely live in the mass range that allows for two different field theory interpretations. On the level of the present discussion, the two different choices simply correspond to an exchange of the role of `source' and `response' function ${\tilde{x}}_{2}(q)$ and ${\tilde{x}}_{5}(q)$~\cite{Klebanov:1999tb}. Accordingly, the two-point function in momentum space is given by \begin{equation} \label{44ex} \langle {\cal{O}}_{44} (0) {\cal{O}}_{44} (q) \rangle \propto q^{\pm6/5} \,, \end{equation} and after Fourier transformation \begin{equation} \label{corr1} \langle {\cal{O}}_{y} (t_{1}) {\cal{O}}_{y} (t_{2}) \rangle \propto \text{TF}^{-1}(q^{\pm6/5})(t_{1}-t_{2}) \propto \frac{1}{|t_{1}-t_{2}|^{1\pm (6/5)}} \,. \end{equation} For the scalars in the {\bf 84}, equation (\ref{scalar}) turns into a Bessel equation (\ref{scan}) with $\lambda=\frac45$, such that its regular solution is given by \begin{equation} \tilde{y}_{(84)}(q,r) = \tilde{r}^{4/5} \, \mathrm{Bessel}_K (2/5,\tilde{r})\,, \end{equation} with near $r=0$ series expansion \begin{equation} \tilde{y}_{(84)}(q,r) = q^{2/5} \, \left( \frac{\Gamma (\frac{2}{5})}{2^{3/5}} \, r^{1/5} + \frac{\Gamma (-\frac{2}{5})}{ 2^{7/5}}\; q^{4/5} \, r^{3/5} + \frac{5 \Gamma (\frac{2}{5})}{12\ 2^{3/5}} \; q^2 \, r^{6/5} + \underset{r \to 0}{o}(r^{6/5}) \right)\,. \end{equation} Thus, the first two coefficients in the expansion (\ref{axy}) are related by \begin{equation} {\tilde{y}}_{3}(q) \propto q^{4/5} \,{\tilde{y}}_{1}(q) \,. \end{equation} Again depending on the choice of assigment $\Delta_{\pm}$, the two-point function is thus given by \begin{equation} \label{corr2} \langle {\cal{O}}_{84} (t_{1}) {\cal{O}}_{84} (t_{2}) \rangle \propto \text{TF}^{-1}(q^{\pm4/5})(t_{1}-t_{2}) \propto \frac{1}{|t_{1}-t_{2}|^{1\pm(4/5)}}\,. \end{equation} \subsection{Comparison to the matrix model}\label{sec:25} The dual field theory is the super matrix quantum mechanics, obtained by dimensional reduction of ten-dimensional SYM theory to one dimension, where it is of the form~\cite{deWit:1988ig} \begin{eqnarray} {\cal L}_{\rm MQM} &=& {\rm tr}\left\{ (D_t \,{\bf X}^k)^2 + \psi^I D_t \psi^I - \frac12[\,{\bf X}^k,\,{\bf X}^l]^2 - \Gamma^k_{IJ}\, \psi^I [\,{\bf X}^k, \psi^J] \right\} \;, \label{sym} \end{eqnarray} with $SU(N)$ valued matrices $\,{\bf X}^k$, $\psi^I$ in the corresponding vector and spinor representations of $SO(9)$. This model itself has been proposed as a non-perturbative definition of M-theory~\cite{Banks:1996vh}. The gauge invariant operators dual to the supergravity scalars in the {\bf 44} and the {\bf 84}, respectively, can be identified via their $SO(9)$ representations \begin{eqnarray} \label{op2} {\cal O}_{44} &\propto& T_{ij}^{++}~=~ \frac{1}{N} \Big( \text{tr} \big( \,{\bf X}^i \,{\bf X}^j \big) - \frac{1}{9}\,\delta^{ij}\, \sum_{k=1}^9 \text{tr} \big( \,{\bf X}^k \,{\bf X}^k \big) \Big)\,, \nonumber\\ {\cal O}_{84} &\propto&J_{ijk} ~\propto~ \frac{1}{N}\, \text{tr} \big( [\,{\bf X}^i,\, \,{\bf X}^j] \,{\bf X}^k \big) \;, \end{eqnarray} The behaviour of these operators in the matrix quantum mechanics has been studied in \cite{Hanada:2011fq} by Monte Carlo methods. Their result shows precise agreement with (\ref{corr1}) and (\ref{corr2}) if we select $\Delta_-$ for the {\bf 44} scalars and $\Delta_+$ for the {\bf 84} scalars, respectively. Only this assignment will correspond to a supersymmetric field theory dual. This result also agrees with the linearized Kaluza-Klein analysis of~\cite{Sekino:1999av} (where the issue of the $\Delta_\pm$ ambiguity was not discussed). In the next section we will use the full non-nonlinear effective theory in order to compute correlation functions for deformations of the model~(\ref{sym}). \section{Deformed BFSS model holography} \label{sec:deformation} In the following section we will construct a half-supersymmetric `deformed' domain-wall solution of two-dimensional $SO(9)$ supergravity which, as it turns out, uplifts to an eleven-dimensional pp-wave with $SO(3)\times SO(6)$ symmetry. We will see however that the resulting eleven-dimensional pp-wave does not belong to the class of bubbling M-theory $SO(3)\times SO(6)$ geometries of \cite{Lin:2004nb}. In particular, contrary to \cite{Lin:2004nb}, our eleven-dimensional pp-wave background has vanishing four-form flux and is consistent with the analysis of \cite{OColgain:2012wv}. From its asymptotic behaviour we conclude that it describes a vev deformation of the BFSS matrix model. In sections \ref{sec:32}, \ref{sec:33} we then use holographic renormalization as developed in the last section to compute around this solution two-point correlation functions of operators dual to the ${\bf 44}$ scalar fields which decompose into \begin{eqnarray} {\bf 44} &\longrightarrow (1,20)\oplus(5,1)\oplus(3,6)\;, \label{44bmn} \end{eqnarray} under $SO(3)\times SO(6)$. \subsection{$SO(3) \times SO(6)$ domain wall} In this section, we determine the half-maximal BPS solutions of the maximal two-dimensional supergravity~(\ref{Ltrunc}) that preserve an $SO(3)\times SO(6)\subset SO(9)$ subgroup of the gauge symmetry. A simple ansatz for such a vacuum solution is provided by exciting the scalars \begin{equation} \label{s3s6trunc} X_{1,2,3} = e^{-x}\;,\quad X_4 = e^{2x}\;, \end{equation} and setting the axion fields $\phi_a$ to zero. The $SO(3)\times SO(6)$ symmetry can be easily seen from the embedding of the $U(1)^4$ truncation (\ref{Ltrunc}) into the full $SO(9)$ theory \cite{Ortiz:2012ib}, where the $SL(9)/SO(9)$ coset space is parametrized by an $SL(9)$ valued scalar matrix ${\cal V}$. In the $U(1)^4$ truncation this matrix is diagonal \begin{equation} {\cal{V}} = \text{diag}\, \big( X_{1}^{-1/2},\,X_{1}^{-1/2},\, \dots ,\, X_{4}^{-1/2},\,X_{4}^{-1/2},\, X_1 X_2 X_3 X_4 \big)\,. \end{equation} With the ansatz \eqref{s3s6trunc}, it takes the form \begin{equation} {\cal{V}}= \begin{pmatrix} e^{x/2} \mathbb{I}_{6\times 6} & 0 \\ 0 & e^{-x} \mathbb{I}_{3\times 3} \\ \end{pmatrix} \,, \label{V36} \end{equation} which preserves an $SO(3)\times SO(6)$ subgroup of the $SO(9)$ gauge symmetry. The two-dimensional bosonic effective Lagrangian (\ref{Ltrunc}) becomes \begin{equation} \label{leff2} {\cal{L}} = -\frac{1}{4} e \rho R + \frac{9}{8} e \rho \, ( \partial_{\mu}x) ( \partial^{\mu} x) + \frac{3}{8} e \rho^{5/9} \, e^{-2x} \, (8 +12 e^{3x} + e^{6x})\,. \end{equation} In the following we will construct BPS solutions in this truncation of the theory. We stress that the $U(1)^4$ truncation (\ref{Ltrunc}) is presumably not the bosonic sector of a supersymmetric theory but can be embedded into the maximally supersymmetric $SO(9)$ theory of \cite{Ortiz:2012ib}, which allows to discuss BPS solutions of the latter. The full theory has 16 gravitinos, 16 dilatinos and 128 fermions. Vanishing of their supersymmetry transformations in the truncation \eqref{s3s6trunc} implies the Killing spinor equations \begin{align} 0 &\overset{!}{=} \partial_{\mu} \epsilon^I + \frac{1}{4} {\omega_{\mu}}^{\alpha \beta} \gamma_{\alpha \beta} \epsilon^I + \frac{7}{12} \,i \rho^{-2/9} (e^{2x}+2e^{-x})\, \gamma_{\mu} \epsilon^I \,,\nonumber\\ 0 &\overset{!}{=} - \frac{i}{2} (\rho^{-1} \partial_{\mu} \rho)\, \gamma^{\mu} \epsilon^I + \frac{3}{4} \rho^{-2/9} (e^{2x}+2e^{-x})\, \epsilon^I\,,\nonumber\\ 0 &\overset{!}{=}(\partial_{\mu} x)\, \gamma^{\mu} \epsilon^I - \frac{2i}{3} \rho^{-2/9} (e^{2x}-e^{-x})\, \epsilon^I\,, \label{KSE} \end{align} for the Killing spinor $\epsilon^I$, $I=1, \dots, 16$\,. Here, $\omega_\mu{}^{\alpha\beta}$ is the spin connection and $\gamma_\alpha$ denote the $SO(1,2)$ gamma matrices. Apart from the $SO(9)$ invariant solution (\ref{dw}) for which $x=0$, these equations admit a unique non-trivial solution. Part of the diffeomorphism invariance can be fixed upon identifying the scalar $x$ with the radial coordinate, after which the solution takes the form \begin{eqnarray} \label{backbmn} \rho(x) &=& e^{\frac{9}{2} x} (e^{3x}-1)^{-9/4}\;,\qquad ds_{2}^2 ~=~ \widetilde{f}(x)^2 dt^2 - \widetilde{g}(x)^2 dx^2\,, \end{eqnarray} with the functions \begin{eqnarray} \widetilde{f}(x) &\equiv& e^{\frac{7}{2} x}(e^{3x}-1)^{-7/4}\,,\qquad \widetilde{g}(x) ~\equiv~\frac{3}{2} e^{2 x} (e^{3x}-1)^{-3/2}\,, \end{eqnarray} up to coordinate redefinitions. The associated Killing spinors are given by \begin{eqnarray} \epsilon^{I}(x) &= a(x) \, \epsilon_{0}^{I}\;,\qquad\mbox{with}\quad \gamma^1 \epsilon_{0}^{I} = -i \epsilon_{0}^{I}\;, \label{KS} \end{eqnarray} and a function $a(x)$ that is obtained from integrating the first equation of (\ref{KSE}). This confirms that the background preserves sixteen supercharges, i.e.\ has the same number of supersymmetries as the $SO(9)$ domain wall (\ref{dw}). Since $x$ is non-vanishing in the bulk, this deformation breaks $SO(9)$ down to $SO(3)\times SO(6)$\,. The Ricci scalar of the two-dimensional metric (\ref{backbmn}) takes the following form \begin{eqnarray} R &=& -\frac{5}{6} \, e^{-2x} \big( e^{6x} -12 e^{3x} - 4 \big)\,, \nonumber\\[1ex] &&{}\mbox{such that}\quad R ~=~ \frac{25}{2} \, + \underset{x \to 0}{{\cal O}}(x^2)\,, \quad R ~=~ -\frac{5}{6} \, e^{4x} + 10\, e^x + \underset{x \to +\infty}{o(1)} \;. \end{eqnarray} It is well defined on the interval $x\in [0\,,\,+\infty[$\, in contrast to the metric and the dilaton which are singular at $x=0$. \paragraph{Higher-dimensional interpretation.} Although the geometry of the solution (\ref{backbmn}) may be obscure in this parametrization, its interpretation becomes clearer in eleven dimensions. Its uplift to ten dimensions can be performed using the embedding of $SO(9)$ supergravity in type IIA supergravity \cite{Anabalon:2013zka}. The resulting solution of the type IIA bosonic equations of motion takes the form \begin{align} ds_{10}^2 &= \rho^{-7/36} \Delta^{7/8} \,ds_2^2 - \rho^{1/4}\,\Delta^{-1/8}\,\Big( \frac{\Delta}{e^x (1-\mu^2)} \, d\mu^2 + e^{-2x}(1-\mu^2) \, d\Omega_{2}^{2} +e^x\mu^2 \, d\Omega_{5}^{2} \Big)\;, \nonumber\\ \Phi&= \frac{1}{3} \log \big( \rho^{-7/4} \Delta^{-9/8}\big) \,,\nonumber\\ F &= 2 \rho^{5/9} \, \big( f_1(x) + \mu^2 \,f_2(x) \big) \, \varepsilon_2 - \frac{3}{2 } \rho\,( *_2 dx) \wedge d(\mu^2)\equiv dA_1\,, \end{align} for metric, dilaton and two-form flux, where \begin{align} 0 &\leq \mu^2 \leq1\,, \quad \Delta \equiv e^{2x}+\mu^2 (e^{-x}-e^{2x})\,,\nonumber\\ f_1 (x)&\equiv -\frac{1}{2} e^{2x} (e^{2x}+6e^{-x}) \,,\quad f_2 (x) \equiv -\frac{1}{2} (e^{-x}-e^{2x}) (4e^{-x}+e^{2x})\,. \end{align} This solution allows straightforward uplift to a purely geometric solution of the $D=11$ Einstein equations according to \begin{align} ds_{11}^{2}&=-2\,dt dz - \frac{\left(e^{3 x}-1\right)^{7/2}}{\left(1-\mu ^2\right) e^{9 x}+\mu ^2 e^{6 x}} \, dz^2 -\frac{1-\mu ^2}{ e^{3 x}-1}\, d\Omega_2^2 -\frac{\mu ^2 e^{3 x}}{ e^{3 x}-1} \, d\Omega_5^2\nonumber\\ &\quad-\frac{9 \, \text{csch}^2\left(\frac{3 x}{2}\right) \left(1-2 \mu ^2+\coth \left(\frac{3 x}{2}\right)\right)}{32} \, dx^2 - \frac{\left(1-\mu ^2\right) e^{3 x}+\mu ^2}{\left(1-\mu ^2\right) \left(e^{3 x}-1\right)} \, d\mu^2\;. \end{align} Eventually, this expression can be considerably simplified by the following coordinate transformations \begin{eqnarray} r_{3}^{2} = \frac{1-\mu ^2}{ e^{3 x}-1}\,,\qquad r_{6}^{2} = \frac{\mu ^2 e^{3 x}}{ e^{3 x}-1}\,,\qquad x^{\pm} = t\pm(t+z)\,, \end{eqnarray} upon which the metric becomes \begin{equation} ds_{11}^{2} = dx^{+} \, dx^{-} - H(r_3,r_6) (dx^{-})^2 - \Big(dr_{3}^{2} + r_{3}^{2}\, d\Omega_2^2 + dr_{6}^{2} + r_{6}^{2}\, d\Omega_5^2 \Big) \,, \label{dw11} \end{equation} where the function $H(r_3,r_6)$ is given by \begin{eqnarray}\label{316} H(r_3,r_6)&\equiv& \frac{(1+\gamma-r_3^2-r_6^2)^{\frac52}(1+\gamma+r_3^2-r_6^2)^{-2}}{\sqrt2\,\gamma\,r_3} \;,\nonumber\\ &&\nonumber\\ \gamma&\equiv& \sqrt{(1 + r_3^2 + r_6^2 + 2 r_6 )(1 + r_3^2 + r_6^2 - 2 r_6 )} \;. \end{eqnarray} Remarkably (but necessarily for consistency) $H(r_3,r_6)$ satisfies the Laplace equation $\Delta H=0$ on the Euclidean space $\mathbb{E}^9$. Consequently the metric (\ref{dw11}) represents a pp-wave solution of eleven-dimensional supergravity \cite{Gauntlett:2002cs}. Just as the domain-wall solution \eqref{2.6}, it is a purely gravitational solution in eleven dimensions. From the ten-dimensional point of view the solution can in fact be interpreted as the near-horizon limit of a distribution of D0 branes with $SO(3)\times SO(6)$ symmetry, similarly to the multi-centered solutions of \cite{Freedman:1999gk} for D3 branes\footnote{We are grateful to the referee of JHEP for bringing up this point.}. To make the form of the distribution explicit, note that $H(r_3,r_6)$ in (\ref{316}) takes the form, \begin{equation} H(r_3,r_6)\sim\frac{1}{r_3}(1-r_6^2)^{-\frac12}+\mathcal{O}(r_3) ~. \end{equation} This suggests that the D0 branes are localized at $r_3=0$ in three of the nine transverse dimensions, while they follow a distribution given by \begin{equation} \sigma(r_6)=\left\{\begin{array}{cr} (1-r_6^2)^{-\frac12}&~,~~ r_6<1\\ 0 &~,~~ r_6\geq 1 \end{array}\right. ~, \end{equation} in the remaining six transverse dimensions. Indeed it can be checked by a direct calculation that \begin{equation} H(r_3,r_6)=\frac{15\sqrt{2}}{2\pi^3}\int\mathrm{d}^9y~\!\delta(\vec{y}_3)~\!\sigma(|\vec{y}_6|)~\!\frac{1}{|\vec{x}_9-\vec{y}_9|^7} ~~\!, \end{equation} where the position vector $\vec{x}_9$ in the transverse directions splits as $\vec{x}_9=\vec{x}_3+\vec{x}_6$ with $r_3:=|\vec{x}_3|$, $r_6:=|\vec{x}_6|$. \paragraph{Operator vs. vev deformation.} Let us consider the 1/2-BPS solution \eqref{backbmn}. After going to the Euclidean signature and making the following Weyl rescaling \begin{equation} g_{\mu \nu} \;\rightarrow\; \rho^{4/9} \, g_{\mu \nu} \;, \label{W49} \end{equation} and coordinate change $(x= r^{2/5})$, one recovers the metric of an asymptotically AdS spacetime coupled to a dilaton: \begin{eqnarray} \label{backW2} d{\widehat{s}_{2}}^2 &=& \widehat{f}(r)^2 dt^2 + \widehat{g}(r)^2 dr^2 \;,\nonumber\\ &&}{ \widehat{g}(r) \equiv \frac{3}{5} x^{-3/2} \, e^{x} (e^{3 x} - 1)^{-1}\,,\quad \widehat{f}(r) \equiv 3^{5/4} \, e^{\frac{5}{2} x} (e^{3 x}-1)^{-5/4}\,. \end{eqnarray} Indeed, up to some global numerical constants, in the limit $(r \to 0)$ one recovers the dilaton coupled AdS background \eqref{backG} \begin{eqnarray} d{\widehat{s}_{2}}^2 &\underset{r \to 0}{\sim}& \frac{dt^2}{r} + \frac{dr^2}{4 r^2} \;, \qquad \rho(t,r) ~\underset{r \to 0}{\sim}~ r^{-9/10}\,. \end{eqnarray} % In this frame where the metric is asymptotically AdS, the near boundary behavior of the scalar field $x(r)$ allows to identify whether the gauge theory dual to the 1/2-BPS solution~\eqref{backbmn} corresponds to an operator deformation or a vev deformation of the undeformed BFSS matrix model~\cite{Klebanov:1999tb,Skenderis:2002wp}. Recall that the correct near-boundary asymptotic form for a scalar $\phi$ propagating in the AdS$_{d+1}$ bulk which is dual to a dimension-$\Delta$ operator in the boundary CFT is given by: \begin{equation} \phi=r^{d-\Delta}\varphi_s+\dots+r^{\Delta}\varphi_v+\dots~\;. \label{asympP} \end{equation} Via the AdS/CFT dictionary $\varphi_s$ is the source for the CFT operator dual to $\phi$, while $\varphi_v$ is its vev (unless the conformal dimension $\Delta$ is in the critical interval which allows for an interchange of the interpretation, as reviewed in appendix~\ref{app:amb}). If instead of an AdS$_{d+1}$ bulk we have an asymptotically AdS$_{d+1}$ geometry which is supported by a nontrivial profile for the bulk field $\phi$ above, we can have two possible scenarios corresponding to two different deformations of the gauge theory: \begin{itemize} \item Operator deformation: this corresponds to an asymptotic behavior $\phi\sim r^{d-\Delta}\varphi_s$ near the boundary. \item Vev deformation: this corresponds to an asymptotic behavior $\phi\sim r^{\Delta}\varphi_s$ near the boundary. \end{itemize} With the general expansion of the active scalar field from (\ref{ax}) \begin{eqnarray} y_{44}(r,t)&= r^{2/5} \, x_{(2)}(t) + r \, x_{(5)}(t) + \dots \;, \label{asyy} \end{eqnarray} we find that around $r=0$, the background \eqref{backW2} \begin{equation} x(r) = r^{2/5}\;, \label{asyx} \end{equation} corresponds to the first term in \eqref{asympP}. However, as we have discussed after (\ref{op2}) above, the BFSS matrix model corresponds to the opposite choice $\Delta_-$ of conformal dimension for the scalar fields in the ${\bf 44}$. I.e.\ the role of source and response in (\ref{asympP}) are exchanged and an asymptotic behavior (\ref{asyx}) of the active scalar field implies the holographic interpretation as a vev deformation. We conclude that the holographic dual to the background \eqref{backbmn} corresponds to a vev deformation of the BFSS model \cite{Skenderis:2002wp}. A domain wall with opposite boundary behaviour on the other hand would describe an operator deformation of the BFSS model such as the BMN matrix model~\cite{Berenstein:2002jq}. The corresponding gravitational background presumably requires also non-vanishing axion fields. In the following, we will compute correlation functions in the deformed matrix model from the gravity side and interpret them in the light of the gauge/gravity correspondence. \subsection{On-shell action and Renormalization} \label{sec:32} The procedure to compute holographic correlation functions around the background \eqref{backbmn} is the same which we have followed in section~\ref{sec:BFSS} for the correlation functions of the BFSS model. As the first step, we will compute the effective action that describes scalar fluctuations around the background \eqref{backbmn}. \subsubsection{Effective action} We will study fluctuations of the full $SO(9)$ supergravity around the background \eqref{backbmn}. To this end we consider the $SL(9)$ valued matrix ${\cal V}$. Its fluctuations are most conveniently expressed by a parametrization \begin{equation} {\cal{V}} \equiv {\cal{V}}_{\text{background}} \, \Big( \mathbb{I}_{9\times9} + X + \frac{1}{2} X^2 + \dots \Big) \;, \end{equation} where ${\cal{V}}_{\text{background}}$ corresponds to the matrix (\ref{V36}) evaluated on the background solution, and $X \in {\mathfrak{sl}}(9)$ carries the scalar fluctuations. Since the background breaks $SO(9)$ down to $SO(3)\times SO(6)$, the fluctuations organize into irreducible representations of $SO(3) \times SO(6)$: \begin{align} {\bf 44}~\longrightarrow~ (1,1) \oplus (5,1) \oplus (1,20) \oplus (3,6) \,. \label{scalarfluc} \end{align} The perturbations $x_{(5,1)}$ and $x_{(1,20)}$ are already captured by the $U(1)^4$ truncation (\ref{s3s6trunc}) and obtained by setting \begin{equation} X_{1,2}=e^{-x+x_{(1,20)}}\,,\quad X_{3}=e^{-x-2x_{(1,20)}} \,,\quad X_4 = e^{2x - 2 x_{(5,1)}}\,. \end{equation} In contrast, the fluctuations in the $(3,6)$ do not sit within the $U(1)^4$ truncation so that their description requires the full $SO(9)$ theory. We will not consider in the following the perturbation in the singlet $(1,1)$, since its interaction with the metric fluctuations leads to rather non-trivial non-diagonal couplings in the action. The resulting Euclidean action quadratic in the scalar fluctuations (\ref{scalarfluc}) is given by \begin{align} S &=- \int dx^2\, e \, \Bigg( -\frac{1}{4} \rho R + \frac{9}{8} e \rho \, ( \partial_{\mu}x) ( \partial^{\mu} x) - \frac{3}{8} e \, \rho^{5/9} \, e^{-2x} (8 +12 e^{3x} + e^{6x})\nonumber\\ &\quad\qquad\qquad+ \frac{1}{2}\,e \rho (\partial x_{(5,1)})^2 + e \, \rho^{5/9} \, e^x (e^{3x}-6) \, x_{(5,1)}^2 \nonumber\\ &\quad\qquad\qquad + \frac{1}{2}\,e \rho (\partial x_{(1,20)})^2 - e \, \rho^{5/9} \,(2 e^{-2x}+ 3 e^x) \, x_{(1,20)}^2 \nonumber\\ &\quad\qquad\qquad +\frac{1}{2}\,e \rho (\partial x_{(3,6)})^2 - e \, \rho^{5/9} \, \frac{e^{-2x}}{2}(3 + 5 e^x + 2 e^{3x}) \, x_{(3,6)}^2 \Bigg)\,. \end{align} As we have seen above, the renormalization process is more easily done after the Weyl rescaling (\ref{W49}) upon which the dilaton enters the action as a global factor. In this frame, the effective action becomes \begin{eqnarray} \label{SE} S &=& \frac{1}{4} \int d^2 x \, e \rho \, \Big( R + \frac{4}{9} \big( \rho^{-1} \partial \rho \big)^2 - \frac{9}{2} \, ( \partial_{\mu}x) ( \partial^{\mu} x) + \frac{3}{2} \, e^{-2x} (8 +12 e^{3x} + e^{6x})\nonumber\\ &&{}\qquad- 2\, (\partial x_{(5,1)})^2 -2\, (\partial x_{(1,20)})^2 -2\, (\partial x_{(3,6)})^2 - 4 \, e^x (e^{3x}-6) \, x_{(5,1)}^2 \nonumber\\ &&{}\qquad + 4 \,(2 e^{-2x}+ 3 e^x) \, x_{(1,20)}^2 + 2 \, e^{-2x}(3 + 5 e^x + 2 e^{3x}) \, x_{(3,6)}^2 \Big)\,. \end{eqnarray} The associated equations of motion are given by \begin{align} 0&=\rho^{-1} \nabla \partial \rho - \frac{3}{2}\, e^{-2 x}(8 +12 e^{3x} + e^{6x}) - \sum_{i\in I} F_{i}(x) \, x_{i}^2 \,,\nonumber\\[1ex] 0&= \rho^{-1} \big(\nabla_{\mu} \partial_{\nu} \rho - \frac{1}{2} g_{\mu \nu} \nabla \partial \rho \big) -\frac{4}{9}\, \rho^{-2} (\partial_{\mu}\rho \partial_{\nu}\rho -\frac{1}{2} g_{\mu \nu} (\partial \rho)^2) +\frac{9}{2}\, \big( \partial_{\mu} x \partial_{\nu} x - \frac{1}{2}\, g_{\mu \nu} (\partial x)^2 \big) \nonumber\\ &\qquad+2 \sum_{i\in I} \big( \partial_{\mu} x_i \partial_{\nu} x_i - \frac{1}{2}\, g_{\mu \nu} (\partial x_i)^2 \big) \,,\nonumber\\[1ex] 0&= R + \frac{4}{9}\, \rho^{-2} (\partial \rho)^2 - \frac{8}{9} \rho^{-1} \nabla \partial \rho - \frac{9}{2} (\partial x)^2 +\frac{3}{2}\, e^{-2 x}(8 +12 e^{3x} + e^{6x}) \nonumber\\ &\qquad-2 \, \sum_{i\in I} \big((\partial x_i)^2- \frac{F_{i}(x)}{2} \, x_{i}^2 \big) \,, \end{align} and \begin{align} 0&= \rho^{-1} \nabla \big(\rho \, \partial x \big) - \frac{2}{3}\, e^{-2 x}(4 -3 e^{3x} - e^{6x}) + \frac{1}{9}\, \sum_{i\in I} F_{i}'(x)\, x_{i}^2 \,,\nonumber\\ 0 &=\rho^{-1}\nabla(\rho \, \partial x_i)+ \frac{1}{2}\, F_{i}(x)\, x_{i}\,, \label{eqmFS} \end{align} with $I \equiv \{(5,1)\,,\,(1,20)\,,\,(3,6) \}$, and the scalar functions \begin{eqnarray} F_{(5,1)} &=& -4 \, e^x (e^{3x}-6) \,,\qquad F_{(1,20)} ~=~ 4 \,(2 e^{-2x}+ 3 e^x) \,, \nonumber\\ F_{(3,6)} &=&2 \, e^{-2x}(3 + 5 e^x + 2 e^{3x}) \,, \end{eqnarray} which capture the interactions of the scalar fluctuations with the background $x(t,r)$ from \eqref{backW2}. \subsubsection{On-shell action and renormalization} Again, the effective action \eqref{SE} is most conveniently evaluated on-shell using the dilaton field equation. As in (\ref{AdSEdW0}) this leads to a contribution located at the boundary of the asymptotically AdS spacetime background \eqref{backW2}, \begin{equation} \label{osaE} S= \frac{1}{2} \int_{r = \epsilon} dt \, \sqrt{|h|} \left(\frac{4}{9} \,n^{\mu} \partial_{\mu} \rho + \rho\, K\right) \,. \end{equation} In the following we will treat the different irreducible representations of the scalar fluctuations separately since they do not mix at the quadratic level. Accordingly, we parametrize the fluctuations of the gravity sector as \begin{align} f(t,r) &= f_{b}(r) \, (1+ f_i(t,r))\,,\nonumber\\ \rho(t,r) &= \rho_{b}(r) \, (1+ \rho_i(t,r))\,, \end{align} where $f_b$ and $\rho_b $ denote the background \eqref{backW2} and the fluctuations \linebreak[0] $\{f_i(t,r),\rho_i(t,r)\}$ are functions of the scalar fluctuations $x_i$ and vanish at the horizon. No source is turned on in the dilaton-gravity sector. % The equations of motion for the scalar fluctuations $x_i$ are given by the last equation of (\ref{eqmFS}) and indicate that a power series expansion in $r$ of the solution should begin with $r^{2/5}$ or $r$, cf.~(\ref{asyy}). Moreover evaluation of the on-shell action \eqref{osaE} on the background shows that the dilaton and extrinsic curvature terms diverge as \begin{align} \sqrt{|h|} \; n^{\mu} \partial_{\mu} \rho &\underset{r \to 0}{\sim} r^{-7/5}\,,\qquad \sqrt{|h|} \; \rho\, K ~\underset{r \to 0}{\sim}~ r^{-7/5}\,. \end{align} Thus we only need to determine the power series expansions up to order $r^{7/5}$, with all the other orders vanishing in the renormalization process. The equations of motion further constrain the expansions to \begin{align} f_i(t,r) &= f_{(4)}(t) \, r^{4/5} + f_{(6)}(t) \, r^{6/5} + f_{(7)}(t) \, r^{7/5} + \dots\,,\nonumber\\ \rho_i(t,r) &= \rho_{(4)}(t) \, r^{4/5} + \rho_{(6)}(t) \, r^{6/5} + \rho_{(7)}(t) \, r^{7/5} + \dots\,,\nonumber\\ x_i (t,r) &= x_{i(2)}(t) \, r^{2/5} + x_{i(4)}(t) \, r^{4/5} + x_{i(5)}(t) \, r + \dots \end{align} Explicitly, the coefficients are related by \begin{align} &f_{(4)}(t)= a_4 \, x_{i(2)}(t)^2 \,, &f_{(6)}(t)&= a_{6} \, x_{i(2)}(t)^2 \,,\nonumber\\ &\rho_{(4)}(t)= b_4 \,x_{i(2)}(t)^2 \,, &\rho_{(6)}(t)&= b_{6} \,x_{i(2)}(t)^2 \,,\nonumber\\ &x_{i(4)}(t) = d_{4} \, x_{i(2)}(t) \,, &\rho_{(7)}(t) &= -\frac{11440}{9} \, x_{i(2)}(t) x_{i(5)}(t) - f_{(7)}(t) \,, \end{align} with the numerical coefficients given by \begin{equation} \begin{array}{c|c|c|c|c|c|} & a_4 & b_4 & d_4 & a_{6} & b_{6} \\ \hline i=(5,1) & -\frac{175}{9} & -35 & -3360 & 847000 & 1524600 \\ \hline i=(1,20) & -\frac{175}{9} & -35 & 4200 & -1001000 & -1801800 \\ \hline i=(3,6) & -\frac{175}{9} & -35 & -12180 & 3003000 & 5405400 \\ \hline \end{array} \end{equation} for the different scalar fields. In particular the coefficients $x_{i(2)}(t)$ and $x_{i(5)}(t)$ are left undetermined in the expansion and should be interpreted as a source and response for the correlation functions. We can now evaluate the on-shell action and renormalize the divergences. The divergences occurring in the on-shell action \eqref{osaE} in the limit $\epsilon \to 0$ are canceled by two counter-terms \begin{align} S_{\text{ct1}} &= \frac{2}{9} \int_{r = \epsilon} dt \, \sqrt{|h|} \, (-\frac92\, \rho -\frac12\, \rho^{1/9}-\frac29\, \rho^{-1/3}) \,,\nonumber\\ S_{\text{ct2}} &= \frac{2}{9} \int_{r = \epsilon} dt \, \sqrt{|h|} \, (\kappa_1\, \rho + \kappa_2\, \rho^{5/9}) \, x_{i}(t,\epsilon)^2 \,, \end{align} which correct the dilaton coupling and the scalar potential, respectively, with the numerical constants given by \begin{equation} \kappa_1~=~ \frac{4}{9} \left(9 a_4 +4 b_4\right) \;,\quad \kappa_2~=~ \frac{2}{27} \left(27\,a_6+a_4\,(9-36\,d_4)+4\,(3\,b_6+b_4-4\,d_4 b_4)\right) \;. \end{equation} Consequently, the renormalized action is given by \begin{align} S_{\text{ren}} &= \lim_{\epsilon \to 0} \left( S_{\text{on-shell}} + S_{\text{ct1}} + S_{\text{ct2}} \right)\nonumber\\ &\propto \quad \int dt \; \big( x_{i (2)}(t) \, x_{i (5)}(t) + \frac{1}{2216}\,\rho_{(7)}(t) \big) \,. \end{align} This expression for the renormalized action is in complete analogy with \eqref{Sren} so in principle one could have guessed the result. Nonetheless, it is interesting to see that the renormalization process developed in \cite{Bianchi:2001de,Bianchi:2001kw,Skenderis:2002wp} straightforwardly works in all cases. In the last step, the coefficients $x_{i(2)}(t)$ and $ x_{i(5)}(t)$ should be related by imposing regularity of the solution in the bulk in order to find the two-point functions by derivation of the action. \subsection{Correlation Functions} \label{sec:33} The computation of correlation functions now proceeds completely in parallel with section~\ref{subsec:correlation}. Let us focus on the scalar two-point functions. They will be generated by the following action \begin{align} S_{\text{ren}} &\propto \int dt \; x_{i (2)}(t) \, x_{i (5)}(t)\nonumber\\ &\propto \int dq \; \tilde{x}_{i (2)}(q) \, \tilde{x}_{i (5)}(q) \,, \end{align} where the functions of the momentum $q$ stand for the coefficients of the Fourier transform of $x_i$. Regularity in the bulk imposes a relation between these two coefficients \begin{equation} \tilde{x}_{i (5)}(q) = C_i (q) \, \tilde{x}_{i (2)}(q) \,, \end{equation} in analogy with (\ref{251}). The two-point function will be given by \begin{equation} \label{cpm} \langle {\cal O}_i(0) {\cal O}_i(q) \rangle \propto C^{\pm1}_i (q) \,, \end{equation} where the plus, minus sign in the exponent should be chosen depending on whether the source is identified with $\tilde{x}_{i (2)}(q)$, $\tilde{x}_{i (5)}(q)$, respectively. In accordance with the discussion of section \ref{sec:25}, the source in the deformed BFSS model should be identified $\tilde{x}_{i (5)}(q)$; this then corresponds to selecting the minus sign in (\ref{cpm}). In the following subsection the function $C_i$ is determined for each scalar perturbation. Unlike for the correlation functions in the undeformed matrix model, we can no longer provide analytical solutions to the scalar fluctuation equations but have to resort to numerical methods to determine the functions $C_i$. \subsubsection{Analytics} The scheme for calculating the two-point functions is now well defined, cf.\ section~\ref{subsec:correlation}.: the first step consists of solving the equations of motion for the scalar perturbations linearized around the background \eqref{backW2}. After taking the Fourier transform with respect to time, we are left with an ordinary second order differential equation in the radial coordinate $r$. There exists a unique solution that is regular in the bulk (i.e.\ falls off sufficiently fast as $r$ goes to infinity). The power series expansion of this regular solution near the horizon $r=0$ allows to compute the ratio \begin{eqnarray} C_i (q)&\equiv& \frac{\tilde{x}_{i (5)}(q)}{\tilde{x}_{i(2)}(q)} \,, \label{ratioC} \end{eqnarray} which describes the two-point function of the dual operators. For computational convenience, we will make the change of variable and field redefinition \begin{equation} u =\sqrt{ e^{3 (r^{2/5})}-1} \,,\qquad \tilde{x}_i (u) \rightarrow u^2 \, \tilde{x}_i (u)\,. \end{equation} The fluctuation equations then translate into \begin{align} 0&= \tilde{x}_{(5,1)}''(u) + \frac{2}{u}\big(\frac{2 u^2 -1}{u^2+1}\big)\,\tilde{x}_{(5,1)}'(u) - \frac{ q^2\,u^3}{(u^2+1)^3}\,\tilde{x}_{(5,1)}(u)\,,\label{flucxxx}\\[1ex] 0&= \tilde{x}_{(1,20)}''(u) + \frac{2}{u}\big(\frac{2 u^2 -1}{u^2+1}\big)\,\tilde{x}_{(1,20)}'(u) + \frac{2u^4- q^2\,u^3-2}{(u^2+1)^3}\,\tilde{x}_{(1,20)}(u)\,,\nonumber\\[1ex] 0&= \tilde{x}_{(3,6)}''(u) + \frac{2}{u}\big(\frac{2 u^2 -1}{u^2+1}\big)\,\tilde{x}_{(3,6)}'(u)\nonumber\\ &\quad+ \frac{2 u^6-q^2 u^5-4 u^4-11 u^2-5+5 (u^2+1)^{1/3} u^2+5u^2(u^2+1)^{1/3}}{u^2\,(u^2+1)^3}\,\tilde{x}_{(3,6)}(u)\,,\nonumber \end{align} for the different species of scalar fields. All solutions admit an expansion \begin{equation} \tilde{x}_i (q,u)= \alpha(q) + \beta(q) \, u^3 + \underset{u \to 0}{{o}}(u^3)\,, \end{equation} at $u=0$ (corresponding to $r=0$), and the ratio (\ref{ratioC}) is given by \begin{eqnarray} C_i &\propto& \frac{\beta(q)}{\alpha(q)} \;. \label{ratioC1} \end{eqnarray} \subsubsection{Numerics} Unlike for the undeformed matrix model, where the regular solution of the scalar fluctuation equations could be found in analytical form (\ref{Bess}), the equations (\ref{flucxxx}) can only be solved numerically. In order to directly extract the ratio (\ref{ratioC1}) of series coefficients in the expansion around $u=0$, we implement a procedure similar to~\cite{Berg:2001ty,Berg:2002hy}. To begin, let us introduce another function \begin{equation} y(q,u) = \tilde{x}(q,u) + \frac{1}{3u} \frac{d\tilde{x}}{du}(q,u)\,, \end{equation} whose power expansion around $u=0$ goes as \begin{equation} y(q,u)= \alpha(q) + \beta(q) \, u + \underset{u \to 0}{{o}}(u^3) \,. \end{equation} For each perturbation, the corresponding equation of motion for $y$ can be solved numerically for given initial conditions at $u=0$. Let $y_1$ and $y_2$ denote the unique solutions with initial conditions \begin{equation} \{\,y_1(0)=1\,,\; y_{1}'(0)=0\,\}\,, \qquad \{\, y_2(0)=0\,,\; y_{2}'(0)=1\,\}\,, \end{equation} respectively, then the unique solution $y_s$ regular in the bulk (when $u\to +\infty$) may be written (up to a global normalization factor) as a linear combination: \begin{equation} y_s = y_1 + \kappa(q) \, y_2 \;= 1 + \kappa(q) \, u + \underset{u \to 0}{{o}}(u^3) = 1 +\frac{\beta(q)}{\alpha(q)}\, u + \underset{u \to 0}{{o}}(u^3)\,. \end{equation} Since $y_1$ and $y_2$ both have the same asymptotic behaviour in the bulk while the combination $y_s$ vanishes, we may read off the quotient $\beta(q)/\alpha(q)$ from the limit \begin{eqnarray} C_i &\propto& \frac{\beta(q)}{\alpha(q)} ~=~ -\lim_{u\rightarrow\infty} \frac{y_1}{y_2} \;, \end{eqnarray} which can be calculated numerically for each value of $q$. A first numerical check suggests that the three ratios \begin{equation} C_{(5,1)}\;,\quad C_{(1,20)}\;,\quad C_{(3,6)}\;, \end{equation} behave like $q^{6/5}$ for large values of $q$. More precisely, for large $q$, these ratios can be fit by a function \begin{equation} C_i ~=~ a_i + b_i\, q^{c_i}\,, \end{equation} with \begin{center} \begin{tabular}{rrr} $a_{(5,1)}=1.72$\,, & $b_{(5,1)}=0.37$\,, & $c_{(5,1)}=1.19$\,, \\ $a_{(1,20)}=1.29$\,, & $b_{(1,20)}=0.37$\,, & $c_{(1,20)}=1.20$\,, \\ $a_{(3,6)}=-18.96$\,, & $b_{(3,6)}=0.80$\,, & $c_{(3,6)}=1.20$\,. \\ \end{tabular} \end{center} In figure~\ref{num}, we have plotted the normalized ratios \begin{equation} r_i(q) \equiv \frac{1}{b_i} \Big( \frac{\tilde{x}_{i\, (5)} (q)}{\tilde{x}_{i\, (2)}(q)} -a_i \Big) \;, \end{equation} in log-log scales, and compared them to the power law $q^{6/5}$ of the undeformed BFSS model \eqref{44ex}. Asymptotically in $q$ we find complete agreement, in accordance with our interpretation of the model as a deformation of BFSS. \begin{figure}[h!] \centering \includegraphics[scale=.8]{BMNasymp4.pdf} \caption{{\small Numerical plot of $C_i$ for the operators dual to the scalar fields (\ref{44bmn}).}} \label{num} \end{figure} \section{Discussion} \label{sec:discussion} We have computed two-point scalar correlation functions in the strong-coupling regime of the BFSS matrix model. The calculation was performed holographically, using as gravitational dual a half-supersymmetric domain wall of the two-dimensional maximally supersymmetric $SO(9)$ gauged supergravity of \cite{Ortiz:2012ib}. This two-dimensional domain wall uplifts to a conformal $AdS_2$ times $S^8$ geometry which is the near horizon limit of $N$ $D0$ branes; a further uplift to eleven dimensions gives an $SO(9)$-symmetric pp-wave. Our results are in agreement with those of~\cite{Sekino:1999av,Sekino:2000mg,Hanada:2011fq}. Furthermore we have constructed a `deformed' half-supersymmetric domain wall which uplifts to an eleven-dimensional pp-wave with broken $SO(3)\times SO(6)$ symmetry. We have argued that this deformation corresponds holographically to turning on an operator vev in the matrix model, and we have used the deformed domain wall as gravitational dual in order to perform a holographic computation of two-point scalar correlation functions. As a consistency check we have verified numerically that in the UV-limit all correlators reduce to those computed in the undeformed BFSS matrix model. This is in accordance with the fact that in the limit of small radial direction the deformed domain-wall solution asymptotes the undeformed domain wall. In principle, similar deformations may exist preserving other maximal subgroups of $SO(9)$. We have chosen $SO(3)\times SO(6)$ since these correspond to the symmetries of the well known BMN operator deformation. However, the corresponding supersymmetric domain wall turned out to be related to a vev rather than to an operator deformation of the BFSS matrix model. Indeed, one may expect that the geometry dual to the BMN matrix model also requires non-vanishing profiles for the axion fields, c.f.~\cite{Lin:2004nb,Lin:2004kw}. The holographic methods of the present paper can be straightforwardly extended to compute matrix model $n$-point functions with $n>2$, which could then in principle be checked independently using Monte Carlo methods directly on the matrix quantum mechanics side. Another possible direction would be the computation of correlation functions in the background of black hole solutions, which corresponds holographically to matrix quantum mechanics at finite temperature. It would also be very interesting to apply these methods to a background which is holographically dual to an operator deformation of the BFSS model, such as the BMN matrix model of~\cite{Berenstein:2002jq}. We plan to return to these questions in the future.
\section{Introduction} A path graph is one of the simplest and most studied interaction topologies of a multi-agent system. In such a system each agent, except for the first and last one, interacts with its two neighbours. Despite its simplicity, a path-graph topology serves, for instance, as a model of vehicular platoons \cite{Tangerman2012, Barooah2009}, discretized flexible structures \cite{Singhose1996} or \cite{Dwivedy2006}, or a spatially-discretized models of long electrical transmission lines \cite{Dhaene1992}. A distributed system defined over a path graph is most often analyzed using state-space techniques and algebraic graph theory. For instance, stability and convergence rate of the system depend on the eigenvalues of the Laplacian matrix \cite{Olfati-Saber2007}. Scaling of $\mathcal{H}_2$ and $\mathcal{H}_\infty$ norms of a path-graph system was investigated in \cite{Herman2014c, Lin2012}. The effect of disturbance was investigated in \cite{Knorn2014} using port-Hamiltonian approach. A good scaling can be captured by a term string stability. Roughly speaking, in a string-stable system a disturbance is not amplified as it propagates among agents (for vehicular platoons \cite{Eyre1998a}, for general systems \cite{Swaroop1996}). However, often the results in the literature rely on the assumption of identical agents. Moreover, they give overall behavior of the system, so it is hard to infer from them what happens in the middle of the platoon or near the boundaries. Following the ideas of other researchers (e.g., \cite{Barooah2009, Galbusera2007}), we will describe the system using wave perspective. Indeed, the propagation of the change in the multi-agent system can be described with the help of travelling waves. We will illustrate it on an example of a system with identical agents and a path-graph topology. If the first agent changes its output, then all following agents sequentially respond to this change. If we study their response from the local point of view \cite{OConnor2007,Martinec2014a,Sirota2015} we can notice that the change is propagated as a wave. The wave departs from the first agent and travels along the system to the last agent, where it reflects and travels back. When it reaches the first agent, it reflects again. The same phenomenon is apparent if the agents are non-identical---the travelling wave is partially reflected on non-identical agents \cite{Martinec2014b}. We can imagine this behaviour as the reflection of the wave if it encounters a boundary between two media of different properties. The tool for analysis will be so called \emph{wave transfer function} (WTF). The transfer-function approach to waves has recently been revisited in a series of papers for lumped models (see \cite{OConnor2007,Yamamoto2013}) and for continuous flexible structures \cite{Sirota2015}. The travelling wave approach has also been applied to vibration control \cite{Mei2011} and it seems to be related to the impedance matching in the power networks \cite{Lesieutre2002}. The wave-based description leads to irrational transfer functions, analysis of which differs in several aspects from their rational counterparts \cite{Curtain2009}. This paper continues in the research started in \cite{Martinec2014a}, where waves in a platoon of identical vehicles is considered. A natural extension of this model is to consider a chain of non-identical (heterogeneous) agents. The first step in the treatment of non-identical agents using travelling waves is given in \cite{OConnor2011} for a mass-spring model. We generalize it by considering an arbitrary dynamics of the agents and their controllers. The preliminary results are presented in \cite{Martinec2014b}, where we introduced the \emph{soft boundary} in a chain of vehicles. Here, we follow this concept of boundaries in a multi-agent system and introduce the second fundamental type of boundary, the \emph{hard boundary}. Although the boundaries are virtual in nature, they principally affect the overall system behaviour. We present some fundamental properties of the boundaries and design wave-absorbing controllers for both types of boundaries. The main contributions of the paper are: i) mathematical description of the travelling waves in a multi-agent system with non-identical agents given by Theorems \ref{soft_boundary_theorem} and \ref{hard_boundary_theorem}, ii) a design of a controller that prevents a reflection of the travelling wave (Theorems \ref{lem:soft_boundary_control} and \ref{lem:hard_boundary_control}) and iii) proof of stability and string stability when these controllers are used (Theorem \ref{lem:soft_boundary_stability}). For better understanding and easy simulations, we provide a set of functions in MATLAB, see \emph{WaveBox} \cite{WaveBox2015}. \vspace{-10pt} \section{System model} We consider a multi-agent system of $N$ non-identical agents with a path-graph interaction topology, for instance, a platoon of vehicles on a highway. In this case the goal of the system is to drive along a line with equal distances between the agents. The dynamics of agents is described by a linear single-input-single-output model with the transfer function $P_i(s)$, where $i$ is the index of the agent. The output, $X_i(s)$, is given as $X_i(s) = P_i(s) U_i(s)$, where $U_i(s)$ is the input to the agent model generated by a control law. This control law is the local relative output feedback between the agent $i$ and its two neighbouring agents $i-1$ and $i+1$. It is modelled as \begin{align} U_i(s) =& C_{\text{f},i}(s)(X_{i-1}(s)-X_{i}(s)) + C_{\text{r},i}(s)(X_{i+1}(s)-X_{i}(s))\nonumber\\ &+ C_{\text{f},i}(s) W_{\text{f},i}(s) + C_{\text{r},i}(s) W_{\text{r},i}(s), \end{align} where $C_{\text{f},i}(s)$ is the transfer function of the controller for the front output difference $X_{i-1}(s)-X_i(s)$ and $C_{\text{r},i}(s)$ is the transfer function of the rear controller. $W_{\text{f},i}(s)$ and $W_{\text{r},i}(s)$ are additional inputs to the agent. We assume that the inputs are equal to zero unless we specify them otherwise. Since in the whole paper we work with transfer functions. we will often omit the argument '$(s)$' when no ambiguity seems possible. Denote the front open-loop transfer function (OLTF) and rear OLTF by $\Mfi{i}(s) = P_{i}(s) C_{\text{f},i}(s)$ and $\Mri{i}(s) = P_{n}(s) C_{\text{r},i}(s)$, respectively. The model of the $i$th agent with zero initial conditions then is \begin{align} X_{i}(s) =& M_{\text{f},i}(s)(X_{i-1}(s)-X_{i}(s)) + M_{\text{r},i}(s)(X_{i+1}(s)-X_{i}(s))\nonumber\\ &+ M_{\text{f},i}(s)W_{\text{f},i}(s) + M_{\text{r},i}(s)W_{\text{r},i}(s).\label{eq:eq1} \end{align} It is depicted in Fig.~\ref{fig:agent_model}. Usually, there is at least one integrator both in the front and rear OLTFs (for instance, from velocity to position), such that the OLTF can be factored as $M(s)=1/s^\nu \overline{M}(s)$, where $\overline{M}(0)<\infty$ and $\nu$ is the number of integrators in the corresponding open loop (either $\Mfi{i}$ or $\Mri{i}$). \begin{figure}[t] \centering \input{agentModel} \caption{The model of $i$th agent.} \label{fig:agent_model} \end{figure} The first agent is described as $X_1(s) = M_{\text{f},1}(s)(W_{\text{f},1}(s)-X_{1}(s)) + M_{\text{r},1}(s)(X_{2}(s)-X_{1}(s))$, where $W_{\text{f},1}(s) = X_{\text{ref}}(s)$ is the external input, which represents the reference for the system (for instance, a virtual leader). This is the input from which the system is to be controlled. The last agent $i=N$ is controlled as $ X_{N}(s) = M_{\text{f},N}(s)(X_{N-1}(s)-X_{N}(s)).$ \subsection{Wave transfer function} The key idea of the wave approach is that the output of the $i$th agent is decomposed into two components, $A_i(s)$ and $B_i(s)$ such that $X_i(s)=A_i(s)+B_i(s)$. The component $A_i(s)$ represents a wave which propagates in the forward direction, that is, to the agents with higher indices. The component $B_i(s)$ represents the wave in the backward direction---to the agents with lower indices. The idea is similar to the standard D'Alambert solution of the wave equation in PDE, where also two waves propagating in different directions appear. Now we summarize the results of \cite{Martinec2014a}, where we considered identical agents. In this case $M_{\text{f},i}(s) = M_{\text{r},i}(s) = M(s)$. The \emph{Wave transfer function} $G(s)$ (abbreviated as WTF) captures how the wave propagates in the system in one direction, that is $A_{i+1}(s)=G(s) A_i(s)$ and $B_{i-1}(s)=G(s) B_i(s)$. There is a simple way how to derive this transfer function. Consider a path graph with infinite number of agents and with only a wave propagating in the forward direction. The wave will never reflect back, so $X_n(s)=A_n(s)$. Then WTF is given by $G(s) = X_{i+1}(s)/X_{i}(s)$ for $N \rightarrow \infty$, see \cite[Sec. 3.1]{Martinec2014a}. When $W_{\text{f},i}\!=\!W_{\text{r},i}\!=\!0$, the system with identical agents is described for $i \in [1, N-1]$ as \cite{Martinec2014a} \begin{align} X_i(s) &= A_i(s) + B_i(s), \label{eq:pos_decomp}\\ A_{i+1}(s) &= G(s)A_i(s), \label{eq:anp1}\\ B_{i}(s) &= G(s)B_{i+1}(s), \label{eq:bnp1}\\ G(s)&= \frac{1}{2}\alpha(s) - \frac{1}{2}\sqrt{\alpha^2(s)-4}, \label{eq:wtf_intro1} \end{align} where $\alpha(s) = 2+ 1/M(s)$, or, alternatively, $\alpha(s) = G(s)+G^{-1}(s)$. The function $G^{-1}(s) = 1/G(s) = \!\frac{1}{2}\alpha(s)\! +\! \frac{1}{2}\sqrt{\alpha^2(s)\!-\!4}$. We now explain the traveling wave concept. Combining (\ref{eq:pos_decomp})-(\ref{eq:bnp1}), \begin{equation} X_i(s) = G(s) A_{i-1}(s) + G(s) B_{i+1}(s). \label{eq:waveProp} \end{equation} This means that the wave $A_{i-1}(s)$ coming from the left is transformed through the transfer function $G(s)$ and summed with the transformed wave from the right $G(s) B_{i+1}(s)$. There are only two boundaries in the homogeneous system. The \emph{forced-end boundary}, located at the first agent, is given b \begin{align} A_1(s) = G(s) W_{\text{f},1}(s) - G^2(s) B_1(s).\label{eq:forced_end} \end{align} and the \emph{free-end boundary}, located at the rear-end agent, is \begin{align} B_N(s) = G(s) A_N(s).\label{eq:free_end} \end{align} The first case is analogous to Dirichlet boundary condition and the second to Neumann. The system with boundaries is shown in Fig.~\ref{fig:homogenous_path_graph}. For both types of the boundaries, a wave-absorbing controller can be designed to prevent a reflection of the incident wave. The controller absorbs the travelling wave by calculating the incident part of the wave and adding this part to its output. With the controller, the input to the first agent changes to \small \begin{align} W_{\text{f},1}(s) &= X_{\text{ref}}(s) +G(s) B_1(s) = G(s) X_1(s) + (1-G(s)^2)X_{\text{ref}}(s).\label{eq:wtf_intro5} \end{align} \normalsize Such absorber can qualitatively improve the transient. This concludes the brief summary of \cite{Martinec2014a}. \begin{figure}[t] \centering \centering \input{wavePropSchematic} \caption{Scheme of a multi-agent system with identical agents. The squares are agents with local dynamics described by (\ref{eq:eq1}). The virtual connections, created by the local control law, are illustrated by springs. } \label{fig:homogenous_path_graph} \end{figure} \subsection{Soft and hard boundaries} Now let us go back to the heterogeneous system of non-identical agents. There are additional boundaries between the agents having different models. In general, we can distinguish between three cases: i) $M_{\text{r},i}(s) \neq M_{\text{f},i+1}(s)$, ii) $M_{\text{f},i}(s) \neq M_{\text{r},i}(s)$, and iii) a combination of i) and ii). As we will see, they cause a partial reflection of the travelling wave. In order to derive its mathematical properties, we assume that the boundary of interest is the only boundary in the system. This also means that all the agents to left of the boundary are identical. Let the WTF for the agents to the left be $G(s)$. Similarly, the agents to the right of the boundary are identical and the WTF for them is $H(s)$. We will focus on the boundaries caused by i) and ii). Once their properties are known, the description of their combination iii) is easy. Thus, we do not discuss it in this paper. \begin{defn} \emph{Soft boundary} is a virtual boundary between two agents, indexed $\sigma$ and $\sigma+1$, with the property $\Mri{\sigma}(s) \neq \Mfi{\sigma+1}(s).$ \end{defn} The soft boundary is, for instance, located in a platoon of non-identical vehicles governed by the same symmetric bidirectional control law or in a mass-spring model with identical springs but non-identical masses. The wave transfer function between the agents with index $i \leq \sigma$ is $G(s)$ and between the agents with $i\geq\sigma+1$ it is $H(s)$, given as \begin{align} G(s) = \frac{1}{2}\alpha_1-\frac{1}{2}\sqrt{\alpha_1^2-4},\;\;\; H(s) = \frac{1}{2}\alpha_2-\sqrt{\alpha_2^2-4}, \label{eq:sb_3} \end{align} where $\alpha_1 = 2 + 1/\Mri{\sigma}(s)$ and $\alpha_2 = 2+1/\Mfi{\sigma+1}(s)$. For the agents with indices $i \leq \sigma$ holds $\Mfi{i}=\Mri{i}=\Mri{\sigma}$ and for agents $i \geq \sigma+1$ holds $\Mfi{i}=\Mri{i}=\Mfi{\sigma+1}$. The irrational transfer functions were obtained identically as in \cite[Sec. 3.1]{Martinec2014a}. The second type of boundary is defined as follows. \begin{defn} The \emph{hard boundary} is a virtual boundary located at the $\eta$th agent with the property $M_{\text{f},\eta} (s) \neq M_{\text{r},\eta} (s).$ \end{defn} The hard boundary is, for instance, located in a platoon of identical vehicles with asymmetric control, see \cite{Barooah2009}, or in a mass-spring model with identical masses but different springs. The adjective `hard' emphasizes the fact that the boundary is located at an agent, in contrast to the soft boundary, located between two agents. To distinguish between the incident, transmitted and reflected waves at the hard boundary, we decompose $X_{\eta}$ to the hard-boundary wave components as \begin{align} X_{\eta}(s) &= A_{\eta,\text{L}}(s) + B_{\eta,\text{L}}(s) = A_{\eta,\text{R}}(s) + B_{\eta,\text{R}}(s), \label{eq:asymmetric2} \end{align} where the indexes $\text{L}$ and $\text{R}$ denote the wave components that are next to the left and right sides of the boundary, respectively. They are tied by the following. \begin{align} A_{\eta,\textnormal{L}}(s) &= G(s) A_{\eta-1}(s), \;\;\; B_{\eta,\textnormal{L}}(s) = G^{-1}(s) B_{\eta-1}(s), \label{eq:asymmetric3}\\ A_{\eta,\textnormal{R}}(s) &= H^{-1}(s) A_{\eta+1}(s), \;\;\; B_{\eta,\textnormal{R}}(s) = H(s) B_{\eta+1}(s), \label{eq:asymmetric4} \end{align} where WTFs to the left and right of the boundary are \begin{align} G(s) = \frac{1}{2}\alpha_1-\frac{1}{2}\sqrt{\alpha_1^2-4},\;\;\; H(s) = \frac{1}{2}\alpha_2-\sqrt{\alpha_2^2-4},\label{eq:asymmetric5} \end{align} and $\alpha_1 = 2 + 1/M_{\textnormal{f},\eta}(s)$ and $\alpha_2 = 2+ 1/M_{\textnormal{r},\eta}(s)$. Both boundaries are illustrated in Fig.~\ref{fig:boundaries_soft_tf}. \begin{figure}[t] \centering \input{softBoundary}\\ \input{hardBoundary} \caption{A system with the soft (top) and hard (bottom) boundary. The blue and red squares are agents with the WTFs $G(s)$ and $H(s)$ in (\ref{eq:sb_3}), (\ref{eq:asymmetric5}). The blue-red spring is the soft boundary and the blue-red square is the hard boundary.} \label{fig:boundaries_soft_tf} \end{figure} \vspace{-10pt} \section{Mathematical description of waves in path graph} In this section we will concentrate on the description of the boundaries and their effect on the propagating wave. \begin{thm}[\cite{Martinec2014b}] \label{soft_boundary_theorem} Let $G(s)$ and $H(s)$ be given by (\ref{eq:sb_3}). Then a soft boundary is in the Laplace domain described by four boundary-transfer functions (BTFs) \begin{alignat}{2} T_{\textnormal{aa}} &= \frac{A_{\sigma+1}}{A_{\sigma}} = \frac{H-H G^2}{1-HG}, \;\;\; T_{\textnormal{ba}}\, &=& \,\frac{A_{\sigma+1}}{B_{\sigma+1}}= \frac{H G - H^2}{1-HG},\label{eq:sb_1}\\ T_{\textnormal{bb}} &= \frac{B_{\sigma}}{B_{\sigma+1}} = \frac{G - H^2 G}{1-HG}, \;\;\; T_{\textnormal{ab}}\, &=&\, \frac{B_{\sigma}}{A_{\sigma}}= \frac{HG - G^2}{1-HG}, \label{eq:sb_2} \end{alignat} \end{thm} The proof is given in \cite{Martinec2014b}. The interpretation is that if there is a wave travelling to the soft boundary from the left-hand side, then it is partially reflected from the boundary (described by $T_{\text{ab}}$) and partially transmitted through the boundary (by $T_{\text{aa}}$). Likewise, if the wave travels from the opposite side, then the transfer functions $T_{\text{ba}}$ and $T_{\text{bb}}$ represent the respective waves. When combined, \begin{align} X_{\sigma}(s) &= G(s) (1+ T_{\text{ab}}(s)) A_{\sigma-1}(s) + T_{\text{bb}}(s) B_{\sigma+1}(s), \label{eq:SB_4a}\\ X_{\sigma+1}(s) &= H(s) (1+ T_{\text{ba}}(s)) B_{\sigma+2}(s) + T_{\text{aa}}(s) A_{\sigma}(s).\label{eq:SB_4b} \end{align} The forced-end boundary is an example of the soft boundary. Substituting $G=0$ into (\ref{eq:sb_1}) and (\ref{eq:sb_2}) gives $T_{\text{aa}} = H$, $T_{\text{ba}} = -H^2$ and $T_{\text{bb}} = T_{\text{ab}} = 0$. In addition, if $G(s)=H(s)$ (there is no boundary), there there is no reflection ($T_\text{ab}(s)=T_\text{ba}(s)=0$) and everything gets through ($T_\text{aa}(s)=G(s), T_\text{bb}(s)=G(s)$). \begin{thm} \label{hard_boundary_theorem} Let $G$ and $H$ be given by (\ref{eq:asymmetric5}). Then the BTFs describing the hard boundary in the Laplace domain are \begin{align} T_{\textnormal{AA}}\! &=\! \frac{A_{\eta,\textnormal{R}}}{A_{\eta,\textnormal{L}}}\!=\!\frac{(1+G)(1-H)}{1-HG}, \;\;\;\; T_{\textnormal{BA}}\!=\! \frac{A_{\eta, \textnormal{R}}}{B_{\eta, \textnormal{R}}}\! =\! \frac{H-G}{1-HG},\label{eq:HB_1a}\\ T_{\textnormal{BB}}\! &=\! \frac{B_{\eta, \textnormal{L}}}{B_{\eta, \textnormal{R}}}\! =\! \frac{(1+H)(1-G)}{1-HG}, \;\;\;\; T_{\textnormal{AB}}\! =\! \frac{B_{\eta, \textnormal{L}}}{A_{\eta, \textnormal{L}}}\! =\! \frac{G-H}{1-HG},\label{eq:HB_1b} \end{align} \end{thm} \begin{pf} From (\ref{eq:eq1}), the output of the agent can be rewritten as \begin{align} X_{\eta}(s) = T_{\text{L}}(s) X_{\eta-1}(s) + T_{\text{R}}(s) X_{\eta+1}(s), \label{eq:app_hard_boundary1} \end{align} where $T_{\text{L}} = M_{\text{f},\eta}/(1+M_{\text{f},\eta} + M_{\text{r},\eta})$ and $T_{\text{R}} = M_{\text{r},\eta}/(1+M_{\text{f},\eta} + M_{\text{r},\eta})$. We combine (\ref{eq:pos_decomp}), (\ref{eq:asymmetric3}) and (\ref{eq:asymmetric4}) to obtain $X_{\eta-1} = A_{\eta-1} + B_{\eta-1} = G^{-1}A_{\eta,\text{L}} + G B_{\eta,\text{L}}$ and $X_{\eta+1} = A_{\eta+1} + B_{\eta+1} = H A_{\eta,\text{R}} + H^{-1} B_{\eta,\text{R}}.$ Substituting these equations into (\ref{eq:app_hard_boundary1}) and using (\ref{eq:asymmetric2}) for $X_{\eta}$ yields \small \begin{align} A_{\eta,\text{L}}(1\!-\!T_{\text{L}} G^{-1})\!+\! B_{\eta,\text{L}}(1\!-\!T_{\text{L}} G)\! &=\! B_{\eta,\text{R}}(T_{\text{R}}H^{-1})\! +\! A_{\eta,\text{R}} T_{\text{R}}H. \label{eq:app_hard_boundary7} \end{align}\normalsize The four wave components are now reduced to three components by substituting $A_{\eta,\text{R}} = A_{\eta,\text{L}} + B_{\eta,\text{L}} - B_{\eta,\text{R}}$ or by $B_{\eta,\text{L}} = A_{\eta,\text{R}} + B_{\eta,\text{R}} - A_{\eta,\text{L}}$ into (\ref{eq:app_hard_boundary7}). These substitutions give \begin{align} B_{\eta,\text{L}} &= B_{\eta,\text{R}} \frac{T_{\text{R}} H -T_{\text{R}} H^{-1}}{T_{\text{L}} G + T_{\text{R}}H-1} + A_{\eta,\text{L}}\frac{1-T_{\text{L}} G^{-1}-T_{\text{R}}H}{T_{\text{L}}G + T_{\text{R}}H-1}, \label{eq:app_hard_boundary8}\\ A_{\eta,\text{R}} &= A_{\eta,\text{L}}\frac{T_{\text{L}}G-T_{\text{L}} G^{-1}}{T_{\text{L}} G + T_{\text{R}} H-1} +B_{\eta,\text{R}}\frac{1-T_{\text{L}} G-T_{\text{R}} H^{-1}}{T_{\text{L}} G + T_{\text{R}} H-1}.\label{eq:app_hard_boundary9} \end{align} These formulas can be further simplified by expressing $T_{\text{L}}$ and $T_{\text{R}}$ in terms of $G$ and $H$. Specifically, $M_{\text{f},\eta} = (G + G^{-1}-2)^{-1}$ and $M_{\text{r},\eta} = (H + H^{-1}-2)^{-1}$. Substituting for them into (\ref{eq:app_hard_boundary8}) and (\ref{eq:app_hard_boundary9}) yields the result. \qed \end{pf} The theorem states that wave incident from the left side of the hard boundary (described by $A_{\eta,\text{L}}$) is partially reflected from the boundary (described by $T_{\text{AB}}$) and partially transmitted through the boundary (by $T_{\text{AA}}$). For the wave incident to the opposite side (described by $B_{\eta,\text{R}}$), the transfer functions are $T_{\text{BA}}$ and $T_{\text{BB}}$, respectively. The output of the hard-boundary agent can be expressed in two equivalent ways, \begin{align} X_{\eta} &= G (1+T_{\text{AB}}) A_{\eta-1} + H T_{\text{BB}} B_{\eta+1},\label{eq:HB_4a}\\ X_{\eta} &= G T_{\text{AA}} A_{\eta-1} + H (1+T_{\text{BA}}) B_{\eta+1}.\label{eq:HB_4b} \end{align} The free-end boundary from (\ref{eq:free_end}) is an example of the hard boundary. In this case $H=0$, which gives $T_{\text{AA}}=T_{\text{BB}}=G$ and $T_{\text{AB}}=T_{\text{BA}} = 0$. Again, if $G(s)=H(s)$, then $T_{AB}=T_{BA}=0$ and $T_{AA}=T_{BB}=1$ which converts (\ref{eq:HB_4a}), (\ref{eq:HB_4b}) into (\ref{eq:pos_decomp}). \subsection{Properties of the boundaries} Although the above descriptions of the boundaries are different, they have some common features. The following result is a direct application of Theorems~\ref{soft_boundary_theorem} and \ref{hard_boundary_theorem}. \begin{cor} \label{cor:HB_tf_relation} The soft and hard BTFs are related as follows, \begin{align} &T_{\textnormal{aa}}(s) = T_{\textnormal{BB}}(s)+G(s)-1,\;\;\ T_{\textnormal{ba}}(s) = H(s) T_{\textnormal{AB}}(s),\\ &T_{\textnormal{bb}}(s) = T_{\textnormal{AA}}(s)+H(s)-1,\;\;\; T_{\textnormal{ab}}(s) = G(s) T_{\textnormal{BA}}(s),\\ &T_{\textnormal{AA}}(s) = 1 + T_{\textnormal{AB}}(s), \;\;\;\;\;\;\;\;\;\;\;\ T_{\textnormal{BB}}(s) = 1 + T_{\textnormal{BA}}(s),\label{eq:HB_tf_rel3}\\ &T_{\textnormal{AA}}(s) + T_{\textnormal{BB}}(s) = 2,\;\;\;\;\;\;\;\;\;\;\;\ T_{\textnormal{AB}}(s) + T_{\textnormal{BA}}(s) = 0.\label{eq:HB_tf_rel5} \end{align} \end{cor} We define the DC gain $\kappa_{\textnormal{G}}$ of $G(s)$ as $\kappa_{\textnormal{G}} = \lim_{s \rightarrow 0} G(s)$. Recall that $\nu$ is the number of integrators in the open loop. \begin{cor} \label{cor:SB_tf_DC_gain} Let $\kappa_{\textnormal{aa}}$ be the DC gain of $T_{\textnormal{aa}}$, $\kappa_{\textnormal{ab}}$ be the DC gain of $T_{\textnormal{ab}}$ etc. If $\nu \geq 1$ for both front and rear OLTFs, then \begin{align} \kappa_{\textnormal{aa}} + \kappa_{\textnormal{bb}} &= 2,\;\;\; \kappa_{\textnormal{ab}} + \kappa_{\textnormal{ba}}= 0, \label{eq:cor_SB_tf_DC_gain2}\\ \kappa_{\textnormal{aa}} - \kappa_{\textnormal{ab}} &= 1,\;\;\; \kappa_{\textnormal{bb}} - \kappa_{\textnormal{ba}} = 1, \label{eq:cor_SB_tf_DC_gain4}\\ \kappa_{\textnormal{aa}} = \kappa_{\textnormal{BB}}, \;\;\; \kappa_{\textnormal{ba}} = &\kappa_{\textnormal{AB}}, \;\;\; \kappa_{\textnormal{bb}} = \kappa_{\textnormal{AA}}, \;\;\; \kappa_{\textnormal{ab}} = \kappa_{\textnormal{BA}}, \end{align} \end{cor} \begin{pf} Under the above assumptions, the DC gain of a WTF is one, i.e. $\lim G(s)_{s \rightarrow 0} = 1$ and $\lim H(s)_{s \rightarrow 0}=1$ \cite{Martinec2014a}. Then, the proof is a straightforward application of Corollary~\ref{cor:HB_tf_relation}. \qed \end{pf} The DC gains of WTFs and BTFs can be used to approximate the value of the transient (see Sec. \ref{sec:local_DC_gains}). The following values of $\kappa_{\text{aa}}$ and $\kappa_{\text{bb}}$ are proved in \ref{appendix_dcgain_sb} \begin{lem} \label{lem:DC_gain_SB} Let ${\numInteg_{\rr}}, {\numInteg_{\ff}}$ be the number of integrators in $M_{\textnormal{r},\sigma}$ and $M_{\textnormal{f},\sigma+1}$, respectively. Assume that ${\numInteg_{\ff}}\geq 1, {\numInteg_{\rr}} \geq 1$. Then we get the DC gains of the soft boundary as \begin{align} \kappa_{\mathrm{aa}} = 2\left(\sqrt{\dfrac{n_{1,0}d_{2,0}}{n_{2,0} d_{1,0}}} +1\right)^{-1}, \, \kappa_{\mathrm{bb}}=2\left(\sqrt{\dfrac{n_{2,0}d_{1,0}}{n_{1,0} d_{2,0}}} +1\right)^{-1}, \end{align} for ${\numInteg_{\ff}}={\numInteg_{\rr}}$, \quad $\kappa_{\mathrm{aa}}=0, \kappa_{\mathrm{bb}}=2$ for ${\numInteg_{\ff}} < {\numInteg_{\rr}},$ \quad and $\kappa_{\mathrm{aa}}=2, \kappa_{\mathrm{bb}}=0$ for ${\numInteg_{\ff}}>{\numInteg_{\rr}}$. The constants are $n_{1,0}/d_{1,0} = \lim_{s \rightarrow 0} s^{\numInteg_{\rr}} M_{\textnormal{r},\sigma}$, $n_{2,0}/d_{2,0} = \lim_{s \rightarrow 0} s^{\numInteg_{\ff}} M_{\textnormal{f},\sigma+1}$. \end{lem} \begin{cor} \label{lem:DC_gains_bounds} If $\nu \geq 1$ both in front OLTF and rear OLTF, then the DC gains are bounded as \begin{align} -1 &\leq \kappa_{\textnormal{ab}}, \kappa_{\textnormal{ba}}, \kappa_{\textnormal{AB}}, \kappa_{\textnormal{BA}} \leq 1,\\ 0 &\leq \kappa_{\textnormal{aa}}, \kappa_{\textnormal{bb}}, \kappa_{\textnormal{AA}}, \kappa_{\textnormal{BB}} \leq 2. \end{align} \end{cor} \begin{pf} The soft-boundary DC gains come from Lemma~\ref{lem:DC_gain_SB} and Corollary~\ref{cor:SB_tf_DC_gain}, the hard-boundary DC gains from Corollary~\ref{cor:SB_tf_DC_gain}.\qed \end{pf} \section{Controllers for the boundaries} We now design a feedback controller compensating the fact that the agents are not identical. The motivation is to prevent the undesired reflection of the wave to shorten the settling time. To control the wave at the boundary, we first have to assume that the input in (\ref{eq:eq1}) is nonzero and investigate its properties. For now, we consider only the input $W_{\text{r},i}$ and we set $W_{\text{f},i}=0$. Hence, the model (\ref{eq:eq1}) of the $n$th agent changes to \begin{align} X_i(s)\!&=\! M_{\text{f},i}(s) (X_{i-1}(s)\!-\!X_i(s))\! +\!M_{\text{r},i}(s) (X_{i+1}(s)\!-\!X_i(s)\!+\!W_{\text{r},i}(s)).\label{eq:add_input_1} \end{align} The input $W_{\text{r},i}(s)$ generates a wave that propagates in the system in the same manner as (\ref{eq:anp1}) and (\ref{eq:bnp1}), only (\ref{eq:waveProp}) is changed as follows \begin{align} X_i(s) = G(s)A_{i-1}(s) + G(s)B_{i+1}(s) + T_{\text{r},i} (s) W_{\text{r},i}(s), \end{align} where $T_{\text{r},i}(s) = X_i(s)/ W_{\text{r},i}(s)$ for $N \rightarrow \infty$. We are interested in finding $T_{\text{r},i}(s)$, hence, we substitute (\ref{eq:anp1})-(\ref{eq:bnp1}) into (\ref{eq:add_input_1}) and get \vspace{-4pt} \begin{align} T_{\text{r},i}(s) = \frac{M_{\text{r},i}(s)}{1 + (1-G(s))M_{\text{f},i}(s) + (1-G(s))M_{\text{r},i}(s)}.\label{eq:add_input_4} \end{align} Analogously, we can calculate the transfer function $T_{\text{f},i}(s) = X_{i}(s)/W_{\text{f},i}$ for a non-zero input $W_{\text{f},i}$. \subsection{The soft boundary controller} \label{subsec:control_soft_boundary} A soft-boundary controller can be designed for various purposes, for instance, to prevent or modify a wave's transmission through the boundary. We now design an absorbing controller that prevents the reflection of a wave from the left side of the soft boundary. The derivation for its right side is analogous. Suppose that the soft boundary is between agents $\sigma$ and $\sigma+1$. By the combination of (\ref{eq:SB_4a}) and (\ref{eq:add_input_4}), we get \begin{align} X_{\sigma} = (1+T_{\text{ab}}) A_{\sigma} + T_{\text{bb}} B_{\sigma+1} + T_{\text{r},i} (1+T_{\text{ab}}) W_{\text{r},\indSB}. \label{eq:control_soft3} \end{align} The reflection of the wave travelling from the left is described by the term $T_{\text{ab}} A_{\sigma}$. Therefore, we set $W_{\text{r},\indSB}(s) = F_{\text{f,S}}(s) A_{\sigma}(s)$, where $F_{\text{f,S}}(s)$ is a transfer function of a controller to be found. To prevent the reflection, we eliminate the term $T_{\text{ab}} A_{\sigma}$ by requiring $T_{\text{ab}}(s)A_{\sigma}(s) + T_{\text{r},i}(s) (1+T_{\text{ab}}(s)) F_{\text{f,S}} (s) A_{\sigma}(s) = 0,$ from which we calculate the controller, \vspace{-7pt} \begin{align} F_{\text{f,S}}(s) = \frac{-T_{\text{ab}}(s)}{T_{\text{r},i}(s)(1+T_{\text{ab}}(s))}. \end{align} Substituting for $T_{\text{ab}}(s)$ from (\ref{eq:sb_2}) and for $T_{\text{r},i}(s)$ from (\ref{eq:add_input_4}), the controller simplifies to $F_{\text{f,S}}(s) = G(s) - H(s).$ It remains to specify $A_{\sigma}$, which represents a wave incident to the soft boundary from the left. By (\ref{eq:pos_decomp})-(\ref{eq:waveProp}), we have $A_{\sigma}(s) = G (X_{\sigma-1} - B_{\sigma-1}) = G X_{\sigma-1} - G^2(X_{\sigma}-A_{\sigma})$, so we can write \vspace{-5pt} \begin{align} A_{\sigma}(s) &= \frac{G(s)}{1-G^2(s)} X_{\sigma-1}(s) - \frac{G^2(s)}{1-G^2(s)} X_{\sigma}(s). \label{eq:control_soft5} \end{align} Therefore, we have everything for the left-side absorbing control law $C_{\text{L,S}}=M_{\text{r},\sigma} F_{\text{f,S}} A_{\sigma}$. Similarly, we can calculate the right-side absorbing control $C_{\text{R,S}}=M_{\text{r},\sigma} F_{\text{r,S}} A_{\sigma}$. They are: \begin{align} C_{\text{L,S}} &= M_{\text{r},\sigma} \frac{G\left( G - H \right)}{1-G^2} \left(X_{\sigma-1} - G X_{\sigma} \right). \label{eq:control_soft6} \\ C_{\textnormal{R},S} &= M_{\textnormal{f},\sigma+1} \frac{H(H-G)}{1-H^2} \left(X_{\sigma+1} - H X_{\sigma} \right).\label{eq:control_soft7} \end{align} \begin{figure}[t] \centering \input{softControllerModel} \caption{The model of $\sigma$th agent with the left-side absorbing controller (blue), where $T_{\text{CLS}} = G(s) (G(s)-H(s))/(1-G^2(s))$ from (\ref{eq:control_soft6}).} \label{fig:soft_controller_model} \end{figure} \begin{thm} \label{lem:soft_boundary_control} The control law preventing any wave to be reflected from the soft boundary in the Laplace domain is \begin{align} X_{\sigma} &= M_{\textnormal{r},\sigma}(X_{\sigma-1} - 2X_{\sigma} +X_{\sigma+1}) + C_{\textnormal{L},S},\label{eq:control_soft11a}\\ X_{\sigma+1} &= M_{\textnormal{f},\sigma+1}(X_{\sigma} - 2X_{\sigma+1} +X_{\sigma+2}) + C_{\textnormal{R},S},\label{eq:control_soft11b} \end{align} where $C_{\textnormal{L},S}, C_{\textnormal{R},S}$ are given in (\ref{eq:control_soft6}), (\ref{eq:control_soft7}). \end{thm} The $\sigma$th agent with implemented left-side absorbing controller is shown in Fig.~\ref{fig:soft_controller_model}. We calculate the wave propagation with the with the left- and right-side absorber implemented. By modifying (\ref{eq:control_soft3}), the outputs of the agents $\sigma$, $\sigma+1$ are \begin{align} X_{\sigma} =& (1+T_{\text{ab}}) A_{\sigma} + T_{\text{bb}} B_{\sigma+1} + F_{\text{f,S}}T_{\text{r},i} (1+T_{\text{ab}}) A_{\sigma} \nonumber\\ &+ T_{\text{bb}} F_{\text{r,S}} T_{\text{f},i} B_{\sigma+1} = G A_{\sigma-1} + G B_{\sigma+1}, \label{eq:control_soft8} \\ X_{\sigma+1} =& H A_{\sigma} + HB_{\sigma+2}.\label{eq:control_soft9} \end{align} Both correspond to a wave propagation in homogeneous system (\ref{eq:waveProp}). Theorem~\ref{lem:soft_boundary_control} describes the only control law that fully absorbs the wave ($T_{\text{ab}} A_\sigma$ in (\ref{eq:control_soft3}) was cancelled by the unique $W_{\text{r},\indSB}$). \subsection{The hard boundary controller} Controlling the hard boundary is similar to that of the soft boundary. Here, we only provide a brief description of the absorbing-controller design. The output of the $\eta$th agent from (\ref{eq:HB_4a}) controlled with additional input $W_{\text{f},\indHB}(s)$ is \begin{align} X_{\eta} = (1 + T_{\text{AB}}) A_{\eta, \text{L}} + T_{\text{BB}}B_{\eta, \text{R}} + T_{\text{f},i} W_{\text{f},\indHB}, \end{align} where $W_{\text{f},\indHB}(s) = F_{\text{f,H}}(s) A_{\eta,\text{L}}(s)$ and $F_{\text{f,H}}(s)$ is a transfer function of a controller that prevents the reflection of a wave. To prevent the reflection of the wave travelling towards the hard boundary from the left, we set $T_{\text{AB}} A_{\eta, \text{L}} =- F_{\text{f,H}} T_{\text{f},i} W_{\text{f},\indHB}$. Hence, \begin{align} F_{\text{f,H}}(s) = -\frac{T_{\text{AB}}(s)}{T_{\text{f},i} (s)} = \frac{(H(s) -G(s))(1-G(s))}{G(s)(1-H(s))}. \end{align} The $A_{\eta, \text{L}}$ term represents the wave travelling towards the hard boundary from left, which is again computed by (\ref{eq:control_soft5}), \begin{align} A_{\eta, \text{L}}(s) = \frac{G(s)}{1-G^2(s)} X_{\eta-1}(s) - \frac{G^2(s)}{1-G^2(s)} X_{\eta}(s). \end{align} The left-side and right-side (calculated similarly) absorbing control laws are then \begin{align} C_{\text{L},\mathrm{H}} &= M_{\text{f},\eta} F_{\text{f,H}} A_{\eta, \text{L}} = M_{\text{f},\eta} \frac{H-G}{(1\!+\!G) (1\!-\!H)}( X_{\eta-1}\! -\! G X_{\eta}), \label{eq:clh} \\ C_{\textnormal{R}, \mathrm{H}} &= M_{\text{r},\eta} F_{\text{r,H}} A_{\eta, \text{R}} = M_{{\mathrm{r}},\eta} \frac{G\!-\!H}{(1\!+\!H) (1\!-\!G)}( X_{\eta+1}\! -\! H X_{\eta}).\label{eq:crh} \end{align} \begin{thm} \label{lem:hard_boundary_control} The control law that prevents any wave to be reflected from the hard boundary in the Laplace domain is \begin{align} X_{\eta} &= M_{\textnormal{f},\eta} (X_{\eta-1}+X_{\eta}) + M_{\textnormal{r},\eta} (X_{\eta+1}-X_{\eta}) + C_{\textnormal{L,H}} + C_{\textnormal{R,H}},\label{eq:control_hard6} \end{align} where $C_{\textnormal{L,H}}, C_{\textnormal{R,H}}$ are given in (\ref{eq:clh}), (\ref{eq:crh}). \end{thm} The control laws implemented on the agent $\eta$ using (\ref{eq:HB_tf_rel3}) give \begin{align} X_{\eta}(s) &= A_{\eta, \text{L}}(s) + B_{\eta, \text{R}}(s) = G(s) A_{\eta-1}(s) + H(s) B_{\eta+1}(s), \label{eq:control_hard9} \end{align} Combining (\ref{eq:control_hard9}) and (\ref{eq:asymmetric2}) gives $A_{\eta, \text{L}} = A_{\eta, \text{R}}$ and $B_{\eta, \text{L}} = B_{\eta, \text{R}}$. In words, the hard boundary between the wave components indexed by $\text{L}$ and $\text{R}$ is removed at the $\eta$th agent and the wave transmits through the agent without being reflected. By comparison of Theorems~{\ref{lem:soft_boundary_control}} and {\ref{lem:hard_boundary_control}} we can see that we need to implement the wave absorber on two agents for the soft boundary but only on one agent for the hard boundary. \begin{figure*}[ht] \centering \includegraphics[width=0.95\textwidth]{comp_soft_bound_N10_J_2_short.pdf} \caption{Simulation of the wave propagating in a multi-agent system with the path-graph topology. A response to a step change in $X_\mathrm{ref}$ is shown. At intermediate times, the wave travels to the soft boundary, where it is transmitted and attenuated by a factor $\kappa_{\text{aa}}$ and reflected by a factor $\kappa_{\text{ab}}$. As the wave propagates back to the first agent, it forces the first four agents to change their output by $\kappa_{\text{ab}}$ (negative value in this case). The waves are absorbed on \nth{1} and \nth{8} agents by the wave-absorbing controllers. The \nth{0} agent is the input to the system from (\ref{eq:wtf_intro5}), i.e. $X_0(s) = W_{\text{f},1}(s)$, $A_0(s) = X_{\text{ref}}(s)$ and $B_0(s) = GB_1(s)$.} \label{fig:comp_soft_bound_N10_J} \end{figure*} \subsection{Stability of the controllers} To prove stability of the control law, we first need to prove stability of the WTF. The proof is given in \ref{appendim_stability_wtf}. \begin{lem} \label{lem:stability_wtf} If $M(s)$ is proper and has no CRHP (closed right half-plane) zeros and no CRHP poles, except for poles at the origin, and if $1+4M(\jmath \omega)$ does not intersect the non-positive real axis for $\omega \in (0,\infty)$, then the wave transfer function $G(s)$ corresponding to $M(s)$ is asymptotically stable. \end{lem} We show below that the controllers do not destabilize the system (proof is in \ref{appendix_soft_boundary_stability}). The opposite is true: their use make the system even string stable for arbitrary platoon size. This guarantees very good transients. \begin{defn}[{\cite{Eyre1998a}}] A system is called $L_2$ string stable if there is an upper bound on the $L_2$-induced system norm of $T_{0,i}(s)$ that does not depend on the number of agents, where $T_{0,i}(s)=\frac{X_i(s)}{X_{\text{ref}}(s)}$.\label{def:string_stability} \end{defn} \begin{thm} \label{lem:soft_boundary_stability} If the WTFs are asymptotically stable, then the multi-agent system with the path-graph topology and the control law from Theorem~\ref{lem:soft_boundary_control} or Theorem~\ref{lem:hard_boundary_control} is asymptotically stable. Furthermore, these control laws implemented with the wave absorber, located on the first or rear-end agent, make the multi-agent system $L_2$ string stable. \end{thm} \subsection{Discussion of wave controllers} The advantage of the wave controller is that it allows the modification of the reflection conditions for the travelling waves on a boundary. Importantly, this modification does not require to change controllers of the other agents in the system and, under certain conditions, it can make the system string stable. Another advantage is that the output of the controller is feasible (see Fig.~\ref{system_input_comp_SB}). A difficulty is that the agent is required to know its own and neighbour's dynamical models. If the neighbour's model is known only approximately, then the wave is not fully absorbed and it partially reflects back. However, the numerical simulations show that the response of the system may still be improved since these schemes are relatively `robust' to the inaccuracies. This is in agreement with experience from practical implementations, see for instance \cite{Saigo2004}, \cite{Kreuzer2011} or \cite{OConnor2008a}. \vspace{-10pt} \section{Numerical simulations of the soft boundary} \label{subsec:Numerical_simulations_SB} The numerical simulations are carried out with \emph{WaveBox}, which is a set of functions and examples in MATLAB that numerically approximates WTFs and BTFs. The \emph{WaveBox} contains a set of examples that show the effect of boundaries and absorbers. Some of the examples are presented in this section. The \emph{WaveBox} was written by the authors and is available at \cite{WaveBox2015}. \subsection{Soft boundary performance} In the numerical simulations we use the following OLTFs, which represent a double integrator agent with viscous friction, controlled by a PI controller: \begin{align} M_1(s) &= \frac{4s+4}{s^2(s+4)}, \quad M_2(s)=\frac{s+k_p}{s^2(s+3)}. \label{eq:agentModelsSim} \end{align} First, consider a path graph of 8 agents with OLTFs $\Mfi{i}(s)=\Mri{i}(s)=M_1(s)$ for $i=1,2,3,4$ and $\Mfi{i}(s)=\Mri{i}(s)=M_2(s)$ for $i=5,6,7,8$ and we set $k_p=1$. There is a soft boundary located between the \nth{4} and \nth{5} agent. The effect of the soft boundary along with wave-absorbing controllers on the first and rear-end agents is demonstrated in Fig.~\ref{fig:comp_soft_bound_N10_J}. The reflection at the boundary is revealed by the negative value of the $B$ component. Since $A$ and $B$ components always sum to to the output, it also validates that our decomposition is correct. Of course, $X_i$ could have been obtained using standard state-space or transfer function approaches, but it would not reveal the local properties (reflections) at the boundary and a way how the signal propagates. This is where our decomposition into waves become useful. Therefore, we think of the wave approach as a complementary tool to the traditional approaches. \begin{figure*}[t] \centering \includegraphics[width=0.95\textwidth]{Six_figs_pos_SB3_short.pdf} \caption{Performance comparison of individual control strategies for a system with soft boundary between agents $4$ and $5$. $x_{\text{ref}}(t) = 1$.} \label{fig:Six_figs_pos_SB} \end{figure*} We have shown in \cite{Martinec2014a} that an absorber implemented on either end shortens the transient in a homogeneous system. In Fig. \ref{fig:Six_figs_pos_SB} we evaluate the performance of absorbers in heterogeneous system described above. Comparing the top-left and bottom-left panels, we can see that the soft-boundary absorber does not shorten the settling time if it is not combined with an absorber on either end. For the absorber on the first agent (top-middle panel), the wave keeps reflecting between the soft boundary and the non-absorbing rear-end agent, which prolongs the transient. The implementation of the soft-boundary absorber (bottom-middle panel) shortens the transient a lot. The absorbers implemented on both the first and rear-end agents (top-right panel) cause a change of the steady-state value, as predicted by Lemma \ref{lem:DC_gain_SB}. This is solved by the soft-boundary absorber (bottom-right panel) or by over-compensating the reference value as we did in \cite{Martinec2014b}. Fig.~\ref{system_input_comp_SB} shows the comparison of the inputs to the fourth agent for (i) the homogeneous system with 8 identical agents (blue solid line) with $\Mfi{i}(s)=\Mri{i}(s)=M_1(s)$, (ii) the system with non-identical agents described above without the soft-boundary absorber (green dashed line), and (iii) the system as in (ii) but with the soft-boundary absorber (red pluses). It illustrates the absorber does not change the control effort too much. \begin{figure}[t] \centering \includegraphics[width=0.49\textwidth]{system_input_comp_SB2.pdf} \caption{The comparison of the inputs to the fourth agent for three different multi-agent systems. The label 'SB abs.' stands for the soft-boundary absorber.} \label{system_input_comp_SB} \end{figure} \subsection{Local effect of the DC gains}\label{sec:local_DC_gains} A local effect of the BTF DC gains is demonstrated in Fig.~\ref{fig:DC_gains_comp}. It shows the first 140 seconds of the step response of system with $N=80$. The models are $\Mfi{i}(s)=\Mri{i}(s)=M_1(s)$ for $i=1,\ldots,40$ and $\Mfi{i}(s)=\Mri{i}(s)=M_2(s)$ for $i=41, \ldots, 80$. The constant $k_p$ is varied. Applying (\ref{eq:pos_decomp})-(\ref{eq:free_end}) and (\ref{eq:SB_4b}), we can describe the output of the \nth{41} agent as \begin{align} X_{41}(s) = T_{\text{aa}}(s) A_{40}(s) + (1+T_{\text{ba}}(s))B_{41}(s), \end{align} where $B_{41}(s)$ is the wave that first transmitted through the boundary, travelled from the \nth{41} agent to the \nth{80} agent, reflected at the end and then travelled back to the \nth{41} agent, hence, $B_{41}(s) = T_{\text{aa}}(s) H^{79}(s) A_{40}(s)$. It takes some time for this reflected wave to propagate, therefore, we can approximate the output $X_{41}(s)$ as $X_{41}(s) \approx T_{\text{aa}}(s) A_{40}(s).$ The approximation is precise in the time-domain until the wave returns back to the \nth{41} agent ($t\approx 110 \,s$). After the initial settling time, the output can be approximated by the DC gain of $T_{\text{aa}}(s)$, because the input is the unit-step in $x_ref(t)=1$. Hence, $x_{41}(t) \approx \kappa_{\text{aa}}, \, t \in [50, 100]$. The most important feature of the approximation is that it does not consider interactions among other agents. In other words, there can be arbitrary number of agents with various dynamics after the \nth{41} agent. \begin{figure}[t] \centering \includegraphics[width=0.49\textwidth]{DC_gains_comp3.pdf} \caption{Output of the $41^{\text{st}}$ agent with $N=140$ for different values of $k_{\text{p}}$.} \label{fig:DC_gains_comp} \end{figure} \section{Conclusion} This paper introduces a wave approach to a multi-agent system with path-graph topology and heterogeneous agents. We define and mathematically describe two basic types of boundaries between non-identical agents. Their effect on the waves travelling in the system is captured using transfer functions, which show that part of the wave gets through the boundary and part is reflected back. The travelling waves describe the system from a local point of view. The wave description allows us to design a feedback controller to compensate the effect of the boundaries, which shortens the settling time. Moreover, such a controller makes the system string stable provided that the system is equipped with at least one wave absorber on the first or rear-end agents.
\section{Introduction} \label{sec:intro} The Casimir effect is a force between conducting plates that arises from the exclusion of vacuum modes by boundary conditions at the plates~\cite{Casimir}. The vacuum energy density between the plates differs from the free density. This defines a separation-dependent effective potential for the plates, and variation with respect to the separaton yields the force. The necessary energy density is obtained from a sum over the allowed modes. The analysis is simplified by working with a massless scalar field and periodic boundary conditions, in place of the complexity of quantum electrodynamics. Contrary to what has been thought, the standard result for the Casimir force between conducting plates at rest in an inertial frame {\em can} be computed in light-front quantization~\cite{LFCasimir}. This is not the same as light-front analyses where the plates are at ``rest'' in an infinite momentum frame~\cite{Lenz}, where the result does not agree. Placement of the plates in the correct frame is critical. The light-front analysis has two important ingredients. One is a careful treatment of the boundary conditions, inspired by the work of Almeida {\em et al}.~\cite{Almeida} on oblique light-front coordinates. The other is the computation of the ordinary energy density, rather than the light-front energy density. The key point here is to focus on the physics; calculations of the same physical effect in different coordinate systems must yield the same result. We define light-front coordinates~\cite{Dirac,DLCQreviews} as $x^+=t+z$ for time and $\ub{x}=(x^-,\vec{x}_\perp)$ for space, with $x^-\equiv t-z$ and $\vec{x}_\perp=(x,y)$. The light-front energy is $p^-=E-p_z$, and the light-front momentum is $\ub{p}=(p^+,\vec{p}_\perp)$, with $p^+\equiv E+p_z$ and $\vec{p}_\perp=(p_x,p_y)$. The mass-shell condition $p^2=m^2$ becomes $p^-=(m^2+p_\perp^2)/p^+$. The natural choice for the analysis of plates at rest in an inertial frame would be equal-time coordinates, with periodicity of the field along one spatial direction. The ``natural'' situation for light-front coordinates is to have spatial periodicity in $x^-$, but this corresponds to plates moving with the speed of light. Lenz and Steinbacher considered this arrangement~\cite{Lenz}. Almeida {\em et al}.~\cite{Almeida} used oblique light-front coordinates with $\bar{x}^0=t+z$ and $\bar{x}^3=z$, and boundary conditions periodic in $z$, to obtain the correct result. Both attempts implied that light-front quantization is deficient in some way; instead, light-front coordinates are just harder to use. In addition to using boundary conditions appropriate to plates at rest in an inertial frame, one must calculate the true vacuum energy. This is not the light-front energy $p^-$. The equal-time energy $E$ is what determines the effective potential and, therefore, the force. Other examples of where this choice matters can be found in the variational analysis of $\phi^4$ theory~\cite{Vary} and the calculation of thermodynamic partition functions. In particular, a partition function should be computed for a system in contact with a heat bath at rest, not at the speed of light~\cite{Elser,SDLCQ}, by use of $e^{-\beta E}$, not $e^{-\beta P^-}$. The standard result for the Casimir force comes from the computation of the expectation value for the energy density, which can be written as a sum over zero-point energies \begin{equation} \label{eq:stdresult} \langle{\cal H}\rangle=\frac{1}{2L}\sum_{n=-\infty}^\infty \int \frac{d^2p_\perp}{(2\pi)^2}E_n, \end{equation} with $E_n=\sqrt{p_\perp^2+\left(\frac{2\pi n}{L}\right)^2}$. The sum can be regulated by a heat-bath factor $e^{-\Lambda E_n}$, to obtain \begin{equation} \langle{\cal H}\rangle=\frac{3}{2\pi^2 \Lambda^4}-\frac{\pi^2}{90 L^4}. \end{equation} The second term is the regulator-independent effective potential and determines the Casimir force. We now consider how this can be done in light-front coordinates. \section{Light-front analysis} \label{sec:analysis} Our goal is to impose periodic boundary conditions on a neutral massless scalar field and compute the vacuum energy density. The light-front mode expansion for the scalar field is \begin{equation} \label{eq:mode} \phi=\int \frac{d\ub{p}}{\sqrt{16\pi^3 p^+}} \left\{ a(\ub{p})e^{-ip\cdot x} + a^\dagger(\ub{p})e^{ip\cdot x}\right\}, \end{equation} with the modes quantized such that \begin{equation} [a(\ub{p}),a^\dagger(\ub{p}')]=\delta(\ub{p}-\ub{p}'). \end{equation} For plates perpendicular to the $z$ axis, placed at $z=0$ and $z=L$, the periodicity imposed is $\phi(z+L)=\phi(z)$. In light-front coordinates, this periodicity is \begin{equation} \phi(x^++L,x^--L,\vec{x}_\perp)=\phi(x^+,x^-,\vec{x}_\perp), \end{equation} which implies $-p^+L/2+p^-L/2=2\pi n$ or $\frac{p_\perp^2}{p^+}-p^+=\frac{4\pi}{L}n$ with $n$ any integer between $-\infty$ and $\infty$. The positive solution of this constraint is \begin{equation} p_n^+\equiv \frac{2\pi}{L}n+\sqrt{\left(\frac{2\pi}{L}n\right)^2+p_\perp^2}. \end{equation} Then $n=-\infty$ corresponds to $p^+=0$, and $n=\infty$ to $p^+=\infty$. A discrete mode expansion can be constructed, with use of discrete annihilation operators \begin{equation} a_n(\vec{p}_\perp)=\sqrt{\left|\frac{dp^+}{dn}\right|}\;a(p_n^+,\vec p_\perp), \end{equation} where $\frac{dp^+}{dn}=\frac{2\pi}{L} \frac{p_n^+}{E_n}$. For these operators, the commutation relation becomes \begin{equation} [a_n(\vec{p}_\perp),a_{n'}^\dagger((\vec{p}_\perp)^{\,\prime})] =\delta_{nn'}\delta(\vec{p}_\perp-\vec{p}_\perp^{\,\prime}). \end{equation} The integration over $p^+$ becomes a sum over $n$: \begin{equation} \int dp^+ =\int \frac{dp^+}{dn}dn\rightarrow \sum_n \frac{dp^+}{dn}. \end{equation} These yield \begin{equation} \phi(x^+=0)=\frac{1}{\sqrt{2L}}\sum_n \int \frac{d^2p_\perp}{2\pi\sqrt{E_n}} \left\{ a_n(\vec{p}_\perp)e^{-ip_n^+x^-/2+i\vec{p}_\perp\cdot\vec{x}_\perp} + a_n^\dagger(\vec{p}_\perp)e^{ip_n^+x^-/2-i\vec{p}_\perp\cdot\vec{x}_\perp}\right\}. \end{equation} The leading $\frac{1}{\sqrt{2L}}$ factor is consistent with normalization on the interval $[-2L,0]$ in $x^-$. For the free scalar, the light-front energy and longitudinal momentum densities are ${\cal H}^-=\frac12|\vec\partial_\perp\phi|^2$ and ${\cal H}^+=2|\partial_-\phi|^2$. Their vacuum expectation values are \begin{eqnarray} \langle0|{\cal H}^-|0\rangle&=&\frac{1}{4L}\sum_{n,n'}\int \frac{d^2p_\perp d^2p'_\perp}{(2\pi)^2\sqrt{E_n E_{n'}}} \vec{p}_\perp\cdot\vec{p}_\perp^{\,\prime} \langle0|a_n(\vec{p}_\perp)a_{n'}^\dagger(\vec{p}_\perp^{\,\prime})|0\rangle \\ &=&\frac{1}{4L}\sum_n \int \frac{d^2p_\perp}{(2\pi)^2 E_n}p_\perp^2, \nonumber \end{eqnarray} \begin{eqnarray} \langle0|{\cal H}^+|0\rangle&=&\frac{2}{2L}\sum_{n,n'}\int \frac{d^2p_\perp d^2p'_\perp}{(2\pi)^2\sqrt{E_n E_{n'}}} \frac{p_n^+ p_{n'}^+}{4} \langle0|a_n(\vec{p}_\perp)a_{n'}^\dagger(\vec{p}_\perp^{\,\prime})|0\rangle \\ &=&\frac{1}{4L}\sum_n \int \frac{d^2p_\perp}{(2\pi)^2 E_n} (p_n^+)^2. \nonumber \end{eqnarray} The energy density, relative to light-cone coordinates, is \begin{eqnarray} {\cal E}_{\rm LF}&\equiv&\frac12(\langle0|{\cal H}^-|0\rangle+\langle0|{\cal H}^+|0\rangle) \\ &=& \frac{1}{8L}\sum_n \int \frac{d^2p_\perp}{(2\pi)^2 E_n} (2E_n^2+2\frac{2\pi}{L}nE_n). \nonumber \end{eqnarray} The second term is zero, because it is proportional to $\sum_{n=-\infty}^\infty n=0$, leaving \begin{equation} {\cal E}_{\rm LF}=\frac{1}{4L}\sum_n\int \frac{d^2p_\perp}{(2\pi)^2}E_n. \end{equation} However, we still need to relate this to the energy density ${\cal E}$ relative to equal-time coordinates. An integration over a finite volume between the plates yields \begin{equation} {\cal E}=\frac{1}{LL_\perp^2}\int_{-2L}^0 dx^- \int_0^{L_\perp} d^2x_\perp {\cal E}_{\rm LF}. \end{equation} A change of variable from $x^-$ to $z=(x^+ + x^-)/2$ at fixed $x^+$ brings \begin{equation} {\cal E}=\frac{1}{LL_\perp^2}\int_0^L 2dx^- \int_0^{L_\perp} d^2x_\perp {\cal E}_{\rm LF}=2{\cal E}_{\rm LF}. \end{equation} Thus, the energy density is \begin{equation} {\cal E}=\frac{1}{2L}\sum_n\int \frac{d^2p_\perp}{(2\pi)^2}E_n, \end{equation} which matches exactly the standard result. It can be regulated with the same heat-bath factor $e^{-\Lambda E_n}$, as appropriate for a system in contact with a heat bath at rest in an inertial frame. For the transverse case, where the plate are separated in a direction transverse to the $z$ axis, the direct implementation of light-front coordinates by Lenz and Steinbacher does yield the correct result. This is not surprising, because plates separated in the transverse direction can be at rest in an inertial frame. However, there could be concern that the additional steps introduced here will somehow destroy this agreement, and we need to check that a consistent result is still obtained. Let the periodicity be in the $x$ direction, with the plates at $x=0$ and $x=L_\perp$, to require $\phi(x^+,x^-,x+L_\perp,y)=\phi(x^+,x^-,x,y)$. The momentum component $p_x$ is then restricted to the discrete values $p_n\equiv 2\pi n/L_\perp$. We define discrete annihilation operators \begin{equation} a_n(p^+,p_y)=\sqrt{\frac{2\pi}{L}}a(p^+,p_n,p_y), \end{equation} with the commutation relation \begin{equation} {[}a_n(p^+,p_y),a_{n'}^\dagger(p^{\prime +},p'_y] =\delta_{nn'} \delta(p^+-p^{\prime +}) \delta(p_y-p'_y). \end{equation} The scalar field is again a discrete sum \begin{eqnarray} \phi(x^+=0)&=&\frac{1}{\sqrt{L_\perp}}\sum_n \int \frac{dp^+ dp_y}{\sqrt{8\pi^2 p^+}} \left\{ a_n(p^+,p_y)e^{-ip^+x^-/2+ip_n x+ip_y y} \right. \\ && \left.+ a_n^\dagger(p^+,p_y)e^{ip^+x^-/2-ip_n x -i p_y y}\right\}. \nonumber \end{eqnarray} The leading factor is consistent with the normalization on the interval $[0,L_\perp]$ in $x$. The energy density is again constructed from the sum of the minus and plus components \begin{eqnarray} \langle0|{\cal H}^-|0\rangle&=&\frac{1}{2L_\perp}\sum_{nn'} \int\frac{dp^+dp_ydp^{\prime+}dp'_y}{8\pi^2\sqrt{p^+p^{\prime+}}} (p_n p_{n'}+p_y p'_y) \\ &&\times \langle0|a_n(p^+,p_y)a_{n'}^\dagger(p^{\prime+},p'_y)|0\rangle \nonumber \\ && =\frac{1}{2L_\perp}\sum_n \int\frac{dp^+ dp_y}{8\pi^2} \frac{p_n^2+p_y^2}{p^+} \nonumber \end{eqnarray} and \begin{eqnarray} \langle0|{\cal H}^+|0\rangle&=& \frac{2}{L_\perp}\sum_{nn'} \int\frac{dp^+dp_ydp^{\prime+}dp'_y}{8\pi^2\sqrt{p^+p^{\prime+}}} \frac{p^+ p^{\prime +}}{4}\langle0|a_n(p^+,p_y)a_{n'}^\dagger(p^{\prime+},p'_y)|0\rangle \\ && =\frac{1}{2L_\perp}\sum_n \int\frac{dp^+ dp_y}{8\pi^2} p^+. \nonumber \end{eqnarray} Averaged together, these fix ${\cal E}_{\rm LF}$ to be \begin{equation} {\cal E}_{\rm LF}=\frac{1}{2L_\perp}\sum_n \int \frac{dp^- dp^+ dp_y}{8\pi^2} \frac{p^- + p^+}{2} \delta\left(p^--\frac{p_n^2+p_y^2}{p^+}\right). \end{equation} The delta function is equivalent to the mass-shell condition \begin{equation} \delta(p^--(p_n^2+p_y^2)/p^+)=p^+\delta(p^2)=p^+\delta(E^2-E_n^2), \end{equation} with $E_n=\sqrt{\left(\frac{2\pi}{L_\perp}n\right)^2 +p_z^2+p_y^2}$. The form of the delta function motivates a conversion to the equal-time variables $E=(p^+ +p^-)/2$ and $p_z=(p^+-p^-)/2$, which yields \begin{equation} {\cal E}_{\rm LF}=\frac{1}{2L_\perp} \sum_n \int \frac{2dE dp_z dp_y}{8\pi^2} E(E+p_z)\frac{1}{2E_n}\delta(E-E_n). \end{equation} The $E$ integral can be done trivially. The contribution from the $p_z$ term is zero, because that part of the $p_z$ integral is odd. This yields the same result as a calculation of the light-front energy density alone; the difference is just the contributions proportional to $p_z$, which integrate to zero. This determines the energy density relative to light-front coordinates as \begin{equation} {\cal E}_{\rm LF}=\frac{1}{4L_\perp}\sum_n\int\frac{dp_z dp_y}{(2\pi)^2}E_n. \end{equation} The transformation to the energy density relative to equal-time coordinates is again just multiplication by two. Therefore, we find in the transverse case \begin{equation} {\cal E}=\frac{1}{2L_\perp}\sum_n\int\frac{dp_z dp_y}{(2\pi)^2} E_n, \end{equation} which matches the usual equal-time result and is of the same form as in the longitudinal case. \section{Summary} \label{sec:summary} We have computed the Casimir effect for parallel plates in light-front coordinates and obtained the standard result, by remaining true to the physics of plates at rest. This is not a simple constraint in light-front coordinates, but is physically correct. Also important for the calculation was the focus on the equal-time energy as the true vacuum energy that determines the effective potential and therefore the Casimir force. This new derivation of the Casimir effect demonstrates that the physics of the effect is independent of the coordinate choice, as it must be, and that light-front quantization is not deficient in its treatment of such vacuum effects. The derivation provides additional confidence in the usefulness and applicability of the light-front approach. \acknowledgments This work was done in collaboration with S.S. Chabysheva and supported in part by the US Department of Energy.
\section{Introduction}\label{sintro} This paper deals with two main topics, the crossing numbers and rectilinear crossing numbers of complete tripartite graphs, and the asymptotic behavior of the crossing number of a balanced complete multipartite graph. In the introduction, we provide background, present definitions, state our main results, and make related conjectures. A plane drawing of a graph is a {\em good drawing} if no more than two edges intersect at any point that is not a vertex, edges incident with a common vertex do not cross, no pair of edges cross more than once, and edges that intersect at a non-vertex must cross. The {\em crossing number} of a good drawing $D$ is the number of non-vertex edge intersections in $D$. The {\em crossing number} of a graph $G$ is \[\crn(G):=\min\{\crn(D) : D \mbox{ is a good drawing of }G\}.\] Clearly a graph $G$ is planar if and only if $\crn(G)=0$. Tur\' an contemplated the question of determining the crossing number of the complete bipartite graph $K_{n,m}$ during World War II, as described in \cite{Turan}. After he posed the problem in lectures in Poland in 1952, Zarankiewicz \cite{Zarank} proved that \[\crn(K_{n,m})\le Z(n,m):=\left\lfloor \frac {n\vphantom{1}} 2\right\rfloor\left\lfloor \frac {n-1} 2\right\rfloor \left\lfloor \frac {m\vphantom{1}} 2\right\rfloor\left\lfloor \frac {m-1} 2\right\rfloor\] and attempted to prove $\crn(K_{n,m})= Z(n,m)$; the latter equality has become known as Zarankiewicz's Conjecture. Hill's Conjecture for the crossing number of the complete graph $K_n$ is \[\crn(K_{n})=H(n):=\frac 1 4 \left\lfloor \frac {n\vphantom{1}} 2\right\rfloor \left\lfloor \frac {n-1} 2\right\rfloor \left\lfloor \frac {n-2} 2\right\rfloor \left\lfloor \frac {n-3} 2\right\rfloor\!,\] and it is known that $\crn(K_{n})\le H(n)$. Background on crossing numbers, including these well-known conjectures, can be found in \cite{BW10} and \cite{RT97MAA}. We establish an upper bound for the rectilinear crossing number of a complete tripartite graph that is analogous to Zarankiewicz' bound. Define \[A(n_1,n_2,n_3) :=\!\!\!\!\!\sum_{i=1,2,3\atop \{j,k\}=\{1,2,3\}\setminus \{i\}}\!\!\!\left( \left\lfloor \frac{n_j\vphantom{1}}{2}\right\rfloor \left\lfloor \frac{n_j-1}{2}\right\rfloor \left\lfloor \frac{n_k\vphantom{1}}{2}\right\rfloor \left\lfloor \frac{n_k-1}{2}\right\rfloor + \left\lfloor \frac{n_i\vphantom{1}}{2}\right\rfloor \left\lfloor \frac{n_i-1}{2}\right\rfloor \left\lfloor \frac{n_j n_k\vphantom{1}}{2}\right\rfloor \right)\!.\] Very little is known about exact values of crossing numbers of complete tripartite graphs, except when two of the parts are small. For example, $\crn(K_{1,3,n})=2\left\lfloor \frac{n}{2}\right\rfloor \left\lfloor \frac{n-1}{2}\right\rfloor + \left\lfloor \frac{n}{2}\right\rfloor$ and $\crn(K_{2,3,n})=4\left\lfloor \frac{n}{2}\right\rfloor \left\lfloor \frac{n-1}{2}\right\rfloor + n$ are established in \cite{A86}, $\crn(K_{1,4,n})=n(n-1)$ is established in \cite{Ho08, HZ08}, and $\crn(K_{2,4,n})=6\left\lfloor \frac{n}{2}\right\rfloor \left\lfloor \frac{n-1}{2}\right\rfloor + 2n$ is established in \cite{Ho13}. It is straightforward to verify that $A(1,3,n)=2\left\lfloor \frac{n}{2}\right\rfloor \left\lfloor \frac{n-1}{2}\right\rfloor + \left\lfloor \frac{n}{2}\right\rfloor=\crn(K_{1,3,n})$, $A(2,3,n)=4\left\lfloor \frac{n}{2}\right\rfloor \left\lfloor \frac{n-1}{2}\right\rfloor + n=\crn(K_{2,3,n})$, $A(1,4,n)=n(n-1)=\crn(K_{1,4,n})$ and $A(2,4,n)=6\left\lfloor \frac{n}{2}\right\rfloor \left\lfloor \frac{n-1}{2}\right\rfloor + 2n=\crn(K_{2,4,n})$. A good planar drawing of $G$ is {\em rectilinear} if every edge is drawn as a straight line segment, and the {\em rectilinear crossing number} $\overline{\operatorname{cr}}(G)$ of $G$ is the minimum number of crossings in a rectilinear drawing of $G$; clearly $\crn(G)\le \overline{\operatorname{cr}}(G)$. Zarankiewicz proved that $\crn(K_{n,m})\le Z(n,m)$ by exhibiting a drawing that actually proves $\overline{\operatorname{cr}}(K_{n,m})\le Z(n,m)$, because the drawing is rectilinear. The next three theorems give bounds on the crossing number and rectilinear crossing number of complete tripartite graphs and are proved in Section \ref{sA3n}. \begin{thm} \label{corAn1n2n3} For all $n_1,n_2,n_3\ge 1$, $\crn(K_{n_1,n_2,n_3})\le\overline{\operatorname{cr}}(K_{n_1,n_2,n_3})\le A(n_1,n_2,n_3)$. \end{thm} \begin{thm}\label{LBthm} For $n$ large enough, $0.666 A(n,n,n) \le \crn(K_{n,n,n})$. \end{thm} \begin{thm}\label{rectLBthm} For $n$ large enough, $0.973 A(n,n,n) \le \overline{\operatorname{cr}}(K_{n,n,n})$. \end{thm} Theorem \ref{LBthm} is proved by a counting argument that has an inherent limitation, whereas Theorem \ref{rectLBthm} is proved by using flag algebras. Theorems \ref{corAn1n2n3}, \ref{LBthm}, and \ref{rectLBthm}, provide evidence for the next two conjectures. \begin{conj} $\overline{\operatorname{cr}}(K_{n_1,n_2,n_3})=A(n_1,n_2,n_3)$. \end{conj} \begin{conj} $\overline{\operatorname{cr}}(K_{n_1,n_2,n_3})=\crn(K_{n_1,n_2,n_3})$. \end{conj} These two conjectures (if true) imply $\crn(K_{n_1,n_2,n_3})=A(n_1,n_2,n_3)$.\medskip A complete multipartite graph is {\em balanced} if the partite sets all have the same cardinality. In \cite{RT97MAA} it is shown that $\lim_{n\to\infty}\frac {\crn(K_n)}{{n\choose 4}}\le \frac 3 8$ and $\lim_{n\to\infty}\frac{\crn(K_{n,n})}{{n\choose 2}^2}\le \frac 1 4$ and the limits exist. We establish an analogous upper bound for the balanced complete $r$-partite graph. The {\em maximum crossing number} of a graph $G$ is \[\CRN(G):=\max\{\crn(D) : D \mbox{ is a good drawing of }G\}.\] With this notation, it is shown in \cite{RT97MAA} that \[\lim_{n\to\infty}\frac {\crn(K_n)}{ \CRN(K_n)}\le \lim_{n\to\infty}\frac{H(n)}{\CRN(K_n)}=\frac 3 8\mbox{ and }\lim_{n\to\infty}\frac {\crn(K_{n,n})}{ \CRN(K_{n,n})}\le \lim_{n\to\infty}\frac{Z(n,n)}{\CRN(K_{n.n})}=\frac 1 4.\] To state our bound for the complete multipartite graph, we need additional notation. The balanced complete $r$-partite graph $K_{n,\dots,n}$ will be denoted by $\Krn r n$ because it is the join of $r$ copies of the complement of ${K_n}$. Note that $\Krn 2 n=K_{n,n}$, $\Krn 3 n=K_{n,n,n}$, and $\Krn n 1=K_n$. \begin{rem} {\rm The maximum crossing number can be computed as the number of choices of $4$ endpoints that can produce a crossing, and can be realized by a rectilinear drawing with vertices evenly spaced on a circle and vertices in the same partite set consecutive (this is well-known for the complete graph and complete bipartite graph). Thus $ \CRN(K_n)={n \choose 4}$ and \begin{equation} \CRN(\Krn r n)={r \choose 2}{n \choose 2}^2+{r}{r-1 \choose 2}{n \choose 2}{n \choose 1}^2+{r \choose 4}{n \choose 1}^4,\label{CReq}\end{equation} with \eqref{CReq} obtained by choosing points partitioned among the partite sets as (2,2), (2,1,1), and (1,1,1,1). For $r=2,3,4$ this yields \begin{enumerate} \item $ \CRN(\Krn 2 n)={n \choose 2}^2$, \item $ \CRN(\Krn 3 n)=3{n \choose 2}^2+3{n \choose 2}{n \choose 1}^2$, \item $ \CRN(\Krn 4 n)=6{n \choose 2}^2+12{n \choose 2}{n \choose 1}^2+{n \choose 1}^4$. \end{enumerate} } \end{rem} A {\em geodesic spherical drawing} of $G$ is a good drawing of $G$ obtained by placing the vertices of $G$ on a sphere, drawing edges as geodesics, and projecting onto the plane. In a {\em random} geodesic drawing, the vertices are placed randomly on the sphere. For integers $r\ge 2$ and $n\ge 1$, define $s(r,n)$ to be the expected number of crossings in a random geodesic spherical drawing of $\Krn r n$ and $\zeta(r):=\frac{3(r^2-r) }{8 \left(r^2+r-3\right)}$. The next theorem is proved in Section \ref{szeta}. \begin{thm}\label{zetathm} For $r\ge 2$, $\lim_{n\to\infty}\frac{s(r,n)}{\CRN(\Krn r n)}=\zeta(r)$. \end{thm} \begin{cor} $\lim\sup_{n\to\infty}\frac{\crn(\Krn r n)}{\CRN(\Krn r n)}\le \zeta(r)$. \end{cor} \begin{obs} Note that $\zeta(r)=\frac{3(r^2-r) }{8 \left(r^2+r-3\right)}$ is monotonically increasing for $r\ge 3$, so $\frac 1 4 =\zeta(2)=\zeta(3)<\zeta(4)<\cdots< \zeta(r)<\zeta(r+1)< \cdots< \frac 3 8$, and $\lim_{r\to\infty}\zeta(r)=\frac 3 8$. \end{obs} \begin{obs} As $n\to\infty$, $\CRN(\Krn 3 n)\approx 3\frac {n^4} 4+ 3 \frac {n^4} 2 =\frac 9 4 n^4$ and $A(n,n, n)\approx 3 \left(\frac {n^4} {16}+\frac {n^4} {8}\right)=\frac 9 {16} n^4$, so $\lim_{n\to\infty} \frac{A(n,n,n)}{\CRN(\Krn 3 n)}=\frac 1 4=\zeta(3)$. \end{obs} \section{Proofs of Theorems \ref{corAn1n2n3}, \ref{LBthm}, and \ref{rectLBthm}}\label{sA3n} In this section we define a drawing of $K_{n_1,n_2,n_3}$ and use it to show that $\overline{\operatorname{cr}}(K_{n_1,n_2,n_3})\le A(n_1,n_2,n_2)$ for all $n_1,n_2,n_3$. We also prove that asymptotically $0.666 A(n,n,n)\le \crn(K_{n,n,n})$ and $0.973 A(n,n,n) \le \overline{\operatorname{cr}}(K_{n,n,n})$ for large $n$. The standard way of producing a rectilinear drawing of the complete bipartite graph $K_{n,m}$ with $Z(n,m)$ crossings is a {\em $2$-line} drawing, constructed by drawing two perpendicular lines and placing the vertices of each partite set on one of the lines, with about half of the points on either side of the intersection of the lines. In the next definition we extend the idea of a 2-line drawing. \begin{figure}[h!] \begin{center} \includegraphics[scale=1.2]{fig-K555} \caption{An alternating 3-line drawing of $K_{5,5,5}$. Points on opposite rays are in one partite set. The partite sets are distinguished by the color of the nodes. The distances were slightly adjusted for visual clarity. The rays and unit circle are shown faint and dotted.} \label{alt3lineK555}\vspace{-15pt} \end{center} \end{figure} \begin{defn}\label{alt3linedef} {\rm An {\em alternating $3$-line drawing} of $K_{n_1,n_2,n_3}$ is produced as follows: \begin{enumerate} \item\label{s1} Draw 3 rays $\overrightarrow{r_1},\overrightarrow{r_2},\overrightarrow{r_3}$ (called the {\em large rays}) that all originate from one point (called the {\em center}) with an angle of $120^\circ$ between each pair of rays. \item For every $i \in \{1,2,3\}$, draw a ray $\overleftarrow{r_i}$ (called a {\em small ray}) from the center in the opposite direction of $\overrightarrow{r_1}$. We call $\overleftarrow{r_i}$ and $\overrightarrow{r_i}$ \emph{opposite} rays, and together they form the $i$th line $\ell_i$. \item\label{spoints} For $i=1,2,3$: \begin{enumerate} \item Define $a_i:=\left\lceil \frac{n_i}{2}\right\rceil$ and $b_i:=\left\lfloor \frac{n_i}{2}\right\rfloor $ \item\label{sa} On $\overrightarrow{r_i}$, place $a_i$ points at distances $\frac 1 {a_i+1}, \frac 2 {a_i+1}, \dots \frac{a_i}{a_i+1}$ from the center. \item\label{sb} On $\overleftarrow{r_i}$, place $b_i$ points at distances $3,4, \dots, b_i+2$ from the center. \end{enumerate} \item\label{s6} For each pair of points not on the same line $\ell_i$, draw the line segment between the points. \end{enumerate} }\end{defn} The rays in Definition \ref{alt3linedef} are not part of the drawing but are useful reference terms. Figure \ref{alt3lineK555} shows an alternating 3-line drawing of $K_{5,5,5}$. The function defined in \eqref{ALeq} below more naturally captures the number of crossings in an alternating 3-line drawing \begin{align}\label{ALeq} A_{3L}(n_1,n_2,n_3):=&\nonumber\\ \sum_{\substack{i=1,2,3\\ \{j,k\}=\{1,2,3\}\setminus \{i\}}}&\left[ {\left\lceil \frac{n_j}{2}\right\rceil \choose 2}{\left\lceil \frac{n_k}{2}\right\rceil \choose 2} +{\left\lceil \frac{n_j}{2}\right\rceil \choose 2}{\left\lfloor \frac{n_k}{2}\right\rfloor \choose 2} + {\left\lfloor \frac{n_j}{2}\right\rfloor \choose 2}{\left\lceil \frac{n_k}{2}\right\rceil \choose 2} + {\left\lfloor \frac{n_j}{2}\right\rfloor \choose 2}{\left\lfloor \frac{n_k}{2}\right\rfloor \choose 2} + \right.\nonumber\\ &\left.\left({\left\lceil \frac{n_i}{2}\right\rceil \choose 2}+{\left\lfloor \frac{n_i}{2}\right\rfloor \choose 2} \right) \left(\left\lfloor \frac{n_j }{2}\right\rfloor\left\lceil \frac{n_k}{2}\right\rceil+\left\lceil \frac{n_j}{2}\right\rceil \left\lfloor \frac{n_k }{2}\right\rfloor \right)\right]. \end{align} \begin{thm}\label{A3nThm} For $n_1,n_2,n_3 \ge 1$, an alternating $3$-line drawing of $K_{n_1,n_2,n_3}$ has at most\break$A_{3L}(n_1,n_2,n_3)$ crossings. \end{thm} \begin{proof} We count the maximum number of possible crossings in an alternating 3-line drawing of $K_{n_1,n_2,n_3}$. There are two types of pairs of points that can result in crossings, (2,2) and (2,1,1), arising from choosing points partitioned among the partite sets as (2,2) and (2,1,1). Throughout this proof, $\{i,j,k\} = \{1,2,3\}$. For type (2,2), if at least one pair of points has a point from each of the large and small ray, i.e., from two opposite rays, we do not get a crossing. Thus we assume each set of two points in the same partite set is actually on the same ray. We can choose any two rays that are not opposite, with each ray to contain two points. There are 12 pairs of rays (omitting the opposite pairs), including 3 cases of $\left\lceil \frac{n_j}{2}\right\rceil$ and $\left\lceil \frac{n_k}{2}\right\rceil$ points, 3 cases of $\left\lfloor \frac{n_j}{2}\right\rfloor $ and $\left\lfloor \frac{n_k}{2}\right\rfloor$ points, and 6 cases of $\left\lceil \frac{n_j}{2}\right\rceil$ points and $\left\lfloor \frac{n_k}{2}\right\rfloor $ points. Thus there are at most \[ \sum_{\substack{i=1,2,3\\ \{j,k\}=\{1,2,3\}\setminus \{i\}}}\left[ {\left\lceil \frac{n_j}{2}\right\rceil \choose 2}{\left\lceil \frac{n_k}{2}\right\rceil \choose 2} +{\left\lceil \frac{n_j}{2}\right\rceil \choose 2}{\left\lfloor \frac{n_k}{2}\right\rfloor \choose 2} + {\left\lfloor \frac{n_j}{2}\right\rfloor \choose 2}{\left\lceil \frac{n_k}{2}\right\rceil \choose 2} + {\left\lfloor \frac{n_j}{2}\right\rfloor \choose 2}{\left\lfloor \frac{n_k}{2}\right\rfloor \choose 2}\right] \] crossings of type (2,2). Consider pairs of points partitioned as type (2,1,1). Denote by $B$ the unit ball centered at the center of the drawing. Observe that the line segment between any two points on small rays is disjoint from $B$ and the line segment between any two points on large rays is entirely in $B$. If we choose the two points in the same partite set from opposite rays, then we do not get a crossing. Thus we assume the two points in the same partite set are actually on the same ray. We can choose any one ray to contain the two points from a (2,1,1) partition of points. Suppose the ray chosen is $\overrightarrow{r_i}$ or $\overleftarrow{r_i}$, where $i \in \{1,2,3\}$. Line $\ell_i$ containing $\overrightarrow{r_i}$ and $\overleftarrow{r_i}$ divides the plane to two half-planes. To have a crossing, the other two points must come from the same half-plane. Thus the number of choices of pairs of points from the two rays in one half plane is $\left\lfloor \frac{n_j }{2}\right\rfloor\left\lceil \frac{n_k}{2}\right\rceil+\left\lceil \frac{n_j}{2}\right\rceil \left\lfloor \frac{n_k }{2}\right\rfloor$. Each of these is multiplied by the choice of pair from $\overrightarrow{r_i}$ and $\overleftarrow{r_i}$, which gives the maximum number of crossings containing a pair from $\overrightarrow{r_i}$ or $\overleftarrow{r_i}$ as \begin{equation} \left({\left\lceil \frac{n_i}{2}\right\rceil \choose 2}+{\left\lfloor \frac{n_i}{2}\right\rfloor \choose 2} \right) \left(\left\lfloor \frac{n_j }{2}\right\rfloor\left\lceil \frac{n_k}{2}\right\rceil+\left\lceil \frac{n_j}{2}\right\rceil \left\lfloor \frac{n_k }{2}\right\rfloor \right).\label{one211}\vspace{-4pt} \end{equation} The maximum number of crossings of type $(2,1,1)$ is obtained by summing \eqref{one211} over all choices of $i \in \{1,2,3\}$. Thus the total number of crossings in this drawing is at most $A_{3L}(n_1,n_2,n_3)$. \ep To complete the proof of Theorem \ref{corAn1n2n3}, we show that $A_{3L}(n_1,n_2,n_3)= A(n_1,n_2,n_3)$. First we show that for all $a$ we have ${ \lceil \frac{a}{2} \rceil \choose 2} + { \lfloor \frac{a}{2} \rfloor \choose 2} = \lfloor \frac{a}{2} \rfloor \lfloor \frac{a-1}{2} \rfloor$. By distinguishing odd and even case we get \[ { \lceil \frac{a}{2} \rceil \choose 2} + { \lfloor \frac{a}{2} \rfloor \choose 2} = \begin{cases} 2{\frac{a}{2} \choose 2} = \frac{a}{2}(\frac{a}{2} -1) = \lfloor \frac{a}{2} \rfloor \lfloor \frac{a-1}{2} \rfloor & \text{ for } a \text { even,} \\ \frac{1}{2} \left( \frac{a+1}{2} \right) \left( \frac{a+1}{2} - 1 \right) + \frac{1}{2} \left( \frac{a-1}{2} \right) \left( \frac{a-1}{2} - 1 \right) = \frac{a^2 - 2a + 1}{4} = \lfloor \frac{a}{2} \rfloor \lfloor \frac{a-1}{2} \rfloor & \text{ for } a \text { odd.} \end{cases} \] Next we show $\lfloor \frac{a}{2} \rfloor \lceil \frac{b}{2} \rceil + \lceil \frac{a}{2} \rceil \lfloor \frac{b}{2} \rfloor = \lfloor \frac{ab}{2} \rfloor$. Again, we distinguish cases by the parity of $a$ and $b$ and obtain \[ \bigg\lfloor \frac{a}{2} \bigg\rfloor \bigg\lceil \frac{b}{2} \bigg\rceil + \bigg\lceil \frac{a}{2} \bigg\rceil \bigg\lfloor \frac{b}{2} \bigg\rfloor = \begin{cases} \frac{a}{2} \left( \lceil \frac{b}{2} \rceil + \lfloor \frac{b}{2} \rfloor \right) = \frac{ab}{2} = \lfloor \frac{ab}{2} \rfloor & \text{ for } a \text{ even},\\% and } b \text{ odd,} \\ \frac{(a-1)(b+1)}{4} + \frac{(a+1)(b-1)}{4} = \frac{2ab -2}{4} = \lfloor \frac{ab}{2} \rfloor & \text{ for } a,b \text{ odd.}\\ \end{cases} \] Using these two observations it is straightforward to show $A(n_1,n_2,n_3)=A_{3L}(n_1,n_2,n_3)$. The assertion that $\overline{\operatorname{cr}}(K_{n_1,n_2,n_3})\le A(n_1,n_2,n_3)$ then follows from Theorem \ref{A3nThm}. This completes the proof of Theorem \ref{corAn1n2n3}.\medskip Next we prove Theorem \ref{LBthm}, i.e., $0.666 A(n,n,n) \le \crn(K_{n,n,n})$ for large $n$. \begin{proof} It is known that $\crn(K_{2,3,n})=4\left\lfloor \frac{n}{2}\right\rfloor \left\lfloor \frac{n-1}{2}\right\rfloor + n$ (see \cite{A86}). So each copy of $K_{2,3,n}$ in $K_{n,n,n}$ has approximately $ n^2$ crossings. The number of copies of $K_{2,3,n}$ in $K_{n,n,n}$ is $6{n\choose 2}{n\choose 3} {n\choose n}$, where the factor of 6 comes from choosing which of the three partite sets in $K_{n,n,n}$ is used for the 2, which for the 3, and which for the $n$. Thus we count about $ \left( n^2\right)\left(6 \cdot \frac {n^2} 2\cdot\frac {n^3} 6 \right)=\frac 1 {2} n^7 $ crossings (counting each crossing multiple times). The number of times a crossing gets counted varies with whether the end points are partitioned of type (2,2) or type (2,1,1). For type (2,2), we can arrange the $K_{2,3,n}$ among the three partite sets as (2,2,0), (2,0,2), or type (0,2,2), and in each case there are 2 choices. Thus a crossing of type (2,2) is counted \begin{eqnarray*} 2\left[{{n-2}\choose{0}}{{n-2}\choose{1}}{{n}\choose{n}} + {{n-2}\choose{0}}{{n}\choose{3}}{{n-2}\choose{n-2}} + {{n}\choose{2}}{{n-2}\choose{1}}{{n-2}\choose{n-2}}\right]& \approx &\\ 2\left[n + \frac{n^3}{6}+ \frac{n^3}{2}\right]&\approx & \frac{4n^4}{3}\vspace{-5pt} \end{eqnarray*} times. For type (2,1,1), we can arrange the $K_{2,3,n}$ among the three partite sets as (2,1,1), (1,2,1), or type (1,1,2), and in each case there are 2 choices. Thus a crossing of type (2,1,1) is counted \begin{eqnarray*} 2\left[{{n-2}\choose{0}}{{n-1}\choose{2}}{{n-1}\choose{n-1}} \!+\! {{n-1}\choose{1}}{{n-2}\choose{1}}{{n-1}\choose{n-1}} \!+\! {{n-1}\choose{1}}{{n-1}\choose{2}}{{n-2}\choose{n-2}}\right]\!\!\!\!& \approx &\!\!\!\!\null\\ 2\left[\frac {n^2} 2 + n^2+ \frac{n^3}{2}\right]&\!\!\approx & \!\!\!\! n^3\vspace{-5pt} \end{eqnarray*} times. Since $\frac{4n^3}{3}>n^3$ and $A(n,n,n)\approx \frac 9 {16} n^4$, asymptotically we have at least \[\frac{\frac 1 2 n^7}{\frac{4n^3}{3}}=\frac 3 8 n^4=\frac 2 3 \left(\frac 9 {16} n^4\right) = \frac 2 3 A(n,n,n)> 0.666 A(n,n,n)\] crossings. \end{proof}\medskip \begin{rem}{\rm We point out that the counting method used in the proof of Theorem \ref{LBthm} has a structural limitation. We use the count number for a (2,2) partition as the number of times a $K_{2,3,n}$ is counted (because it is the larger), even though we know that asymptotically 2/3 of the crossings in an alternating 3-line drawing of $K_{n,n,n}$ are of type (2,1,1) rather than (2,2). So even with the assumption that $\crn(K_{n,n,n})=A(n,n,n)$, this method cannot be expected to produce a lower bound of $cA(n,n,n)$ with $c$ close to 1. }\end{rem} Finally we prove Theorem \ref{rectLBthm}, i.e., $0.973 A(n,n,n) \le \overline{\operatorname{cr}}(K_{n,n,n})$ for large $n$. The proof uses flag algebras, a method developed by Razborov~\cite{Raz07}. A brief explanation of this technique specific to its use in our proof can also be found in Appendix \ref{appdx1}. We use an approach similar to the technique Norin and Zwols~\cite{NorinZwols} used to show that $ 0.905 Z(m,n) \leq \crn(K_{m,n})$; however, we restrict our attention to rectilinear drawings. \begin{proof} For sufficiently large $n$, we first use flag algebra to methods show that in any rectilinear drawing of $K_{n,n,n}$ the average number of crossings over all the copies of $K_{3,2,2}$ that appear in $K_{n,n,n}$ is greater than 5.6767. In our application, we record in flags crossings and tripartitions. We ignore rest of the embedding. Let $G$ be a tripartite graph on $n$ vertices with a rectilinear drawing. A corresponding flag $F_G$ on $n$ vertices $V$ contains a function $\varrho_1: V \rightarrow \{0,1,2\}$, which records the partition of the vertices, and a function $\varrho_2: V^4 \rightarrow \{0,1\}$, which record crossings. We define $\varrho_2(a_1,a_2,b_1,b_2) = 1$ if the vertices of $G$ corresponding to $a_1$ and $a_2$ form an edge of $G$ that crosses an edge of $G$ formed by $b_1$ and $b_2$, and 0 otherwise. We use flags on 7 vertices obtained from rectilinear drawings of $K_{3,2,2}$ (so $m=7$ in Equation \eqref{eq:sum} in Appendix \ref{appdx1}). All rectilinear drawings of $K_7$ were obtained by Aichholzer, Aurenhammer and Krasser~\cite{AichholzerAK:2002}. The drawings give us 6595 flags. We generate 42 types, which leads to 42 equations like \eqref{eq:theta} in Appendix \ref{appdx1}. The optimal linear combination of these equations is computed by CSDP~\cite{Borchers:1999}, an open source semidefinite program solver. CSDP is a numerical solver that provides a positive semidefinite matrix $M$ of floating point numbers. We round the matrix $M$ in Sage~\cite{sw:sage} to a positive semidefinite matrix $Q$ with rational entries. The rounding is done by decomposing $M = U^T D U$ (where $D$ is a diagonal matrix of eigenvalues and $U$ is a real orthogonal matrix of eigenvectors), rounding the entries of $D$ and $U$ to rational matrices $\hat D$ and $\hat U$, and constructing matrix $Q=\hat U^T \hat D\hat U$. Then we use $Q$ to compute the resulting bound $1419186177261/250000000000 > 5.6767$. Software needed to perform the whole computation is available at \url{https://orion.math.iastate.edu/lidicky/pub/knnn}. In a complete graph on 7 vertices, the number of 4-tuples of points is ${7 \choose 4}=35$. Thus the `density' of crossings in $K_{3,2,2}$ is at least $\frac {5.6767} {35}$. The graph $K_{n,n,n}$ must have at least this density times the number of 4-tuples, and the number of 4-tuples is ${3n\choose 4}\approx \frac {81n^4}{24}$. Since $A(n,n,n)\approx \frac {9n^4}{16}$, asymptotically \[\overline{\operatorname{cr}}(K_{n,n,n})> \left(\frac {5.6767} {35}\right)\left(\frac {81n^4}{24}\right) \approx \left(\frac {5.6767} {35}\right)6A(n,n,n)>.973A(n,n,n).\qedhere\] \end{proof}\medskip \begin{rem} {\rm The flag algebra method just applied to $\overline{\operatorname{cr}}(K_{n,n,n})$ with $n\to\infty$ will also work for $\overline{\operatorname{cr}}(K_{n_1,n_2,n_3})$ where $n_i\to\infty$ for all $i=1,2,3$.}\end{rem} \section{Proof of Theorem \ref{zetathm}}\label{szeta} We need a preliminary lemma. \begin{lem}\label{paircrossprob} In a random geodesic spherical drawing of a pair of disjoint edges, the probability that the pair crosses is $\frac 1 8$. \end{lem} \begin{proof} A pair of edges is determined by two sets of endpoints. Each set of two endpoints determines a great circle, and these two great circles intersect in two antipodal points. These two antipodal points of intersection are the potential crossing points, and a crossing occurs if and only if both edges include the same antipodal point. Notice that first picking two great circles uniformly at random, and then picking two points uniformly at random from each of the great circles is equivalent to picking two pairs of points uniformly at random from the sphere. Therefore, for each set of two endpoints, the probability that the great circle geodesic between them includes one of the two antipodal points is $\frac 1 2$, so the probability that both edges include an antipodal point is $\frac 1 4$. Half the time these are the same antipodal point. \end{proof}\medskip We are now ready to prove Theorem \ref{zetathm}, i.e., $\lim_{n\to\infty}\frac{s(r,n)}{\CRN(\Krn r n)}=\zeta(r)$. \begin{proof} The probability of getting a crossing among four points in a geodesic spherical drawing of $\Krn r n$ depends on how the points are partitioned among the partite sets, because different partitions of four points have different numbers of pairs of disjoint edges. Define three types of partitions of four points, classified by the number of pairs (of disjoint edges) produced. \noindent{\bf Type A} 0 pairs: The four points are partitioned among partite sets as (4) or (3,1). Let $\alpha_r$ denote the probability that four randomly chosen points in $\Krn r n$ are of this type. \noindent{\bf Type B} 2 pairs: The four points are partitioned among partite sets as (2,2) or (2,1,1). Let $\beta_r$ denote the probability that four randomly chosen points in $\Krn r n$ are of this type. \noindent{\bf Type C} 3 pairs: The four points are partitioned among partite sets as (1,1,1,1). Let $\gamma_r$ denote the probability that four randomly chosen points in $\Krn r n$ are of this type. We assume that $n$ is large relative to $r$, so we can ignore the difference between $n-1$ and $n$, etc., and we focus only on which partite sets are chosen. For Type C we must choose four distinct partite sets, so $\gamma_r=\frac {r(r-1)(r-2)(r-3)}{r^4}=\frac {(r-1)(r-2)(r-3)}{r^3}$. For Type A there are two choices. For partition (4) the probability is $\frac 1 {r^3}$. To determine the probability of partition (3,1) we count the ways that we can choose 4 partite sets with 3 of them being the same set (which we call a (3,1) choice), and divide by $r^4=$ the number of all possible arrangements of four points into $r$ partite sets. A (3,1) choice can be made by first choosing two distinct partite sets (there are $r(r-1)$ ways to select the two, with the first choice to appear 3 times) and then indicating the order of these partite sets (there are four different orders, determined by where the singleton is placed in the order). So the probability is $\frac {4r(r-1)} {r^4}=\frac {4(r-1)} {r^3}$. Thus $\alpha_r=\frac {4(r-1)} {r^3}+\frac {1} {r^3}=\frac {4r-3} {r^3}$. Then $\beta_r=1-\alpha_r-\gamma_r$. Let $q$ be the number of 4-tuples of points. By Lemma \ref{paircrossprob}, the expected number of crossings in a geodesic spherical drawing is $\frac 1 8$ the number of pairs of disjoint edges, and the number of pairs is \[(3\gamma_r+2\beta_r)q=(3\gamma_r+2(1-\alpha_r-\gamma_r))q=(2+\gamma_r-2\alpha_r)q,\] so $s(r,n)=\frac 1 8 (2+\gamma_r-2\alpha_r)q$. In the earlier described drawing that maximizes the number of crossings, every 4-tuple of Type B and C produces one crossing. There are $(\beta_r+\gamma_r)q$ such 4-tuples, and therefore \[\CRN\left(\Krn r n\right)=(\gamma_r+\beta_r) =(1-\alpha_r)q.\] Thus \begin{eqnarray*} \lim_{n\to\infty}\frac{s(r,n)}{\CRN(\Krn r n)}&=& \frac{\frac 1 8 (2+\gamma_r-2\alpha_r)q}{(1-\alpha_r)q}\\ &=&\frac{2 r^3 + (r - 1) (r - 2) (r - 3) - 2 (4 r - 3)}{8 \left(r^3 - (4 r - 3)\right)}\\ &=&\frac{3(r^2-r) }{8 \left(r^2+r-3\right)}\\ &=&\zeta(r).\hspace{6cm}\qedhere\end{eqnarray*} \end{proof}\medskip \subsection*{Acknowledgments} This research began at the American Institute of Mathematics workshop Exact Crossing Numbers, and the authors thank AIM. The authors thank Sergey Norin for many helpful conversations during that workshop. This paper was finished while Hogben, Lidick\'y, and Young were general members in residence at the Institute for Mathematics and its Applications, and they thank IMA. The authors also thank NSF for their support of these institutes. The work of Gethner and Pfender is supported in part by their respective Simons Foundation Collaboration Grants for Mathematicians. The work of Lidick\'y is partially supported by NSF grant DMS-12660166.
\section{\label{sec:level1}Introduction} Superfluidity (SF), i.e., the emergence of frictionless flow below a critical temperature, is one of the most astonishing manifestations of quantum coherence on a macroscopic scale. It was first observed in 1938 in liquid 4-He. In bulk systems, SF is known to being closely related to Bose-Einstein condensation (BEC) and a consequence of off-diagonal long range order of the density matrix.\citep{yushi} This concept, obviously, is not directly applicable to finite size systems in traps. Moroever, these systems are possibly strongly inhomogeneous, as a result of boundary or confinement effects. An alternative approach to superfluidity is then to use the probability of realization of exchange cycles involving a sufficiently large number of particles. A well studied example for such mesoscopic quantum systems are parahydrogen clusters where, for a small number of particles, $N\sim 10$, the superfluid properties explicitly depend on the precise particle number and the symmetry,\citep{khai} and the superfluidity is essentially uniformly distributed among the system.\citep{mezza,idowu} In addition, both solid-like order and SF do coexist,\citep{idowu} thus qualifying as a supersolid (see e.g. \citep{supersolid}) state of matter. Another interesting test case are harmonically confined two-dimensional Coulomb clusters, whose behavior can be controlled by varying the confinement field strength. There it has been found that not only the global superfluid fraction, but also the spatial distribution of superfluidity crucially depends on $N$.\citep{fili} In particular, for magic numbers, e.g. $N=12,19,\dots$ (closed shells), there is a strong hexagonal symmetry at the center of the trap and the superfluid density is pushed to the boundary of the system. While the dependence of superfluidity on temperature, system size and coupling strength is well understood, the effect of the system dimensionality in finite systems has not been studied systematically. In this work, we try to (partly) fill this gap by studying large spherically confined Coulomb clusters in $2D$ and $3D$ and compare results for, both, the global and spatial distribution of SF. The selected system size $N=150$ is representative and large enough to eliminate the sensitivity of the properties to the precise particle number.\cite{ludwig05} Furthermore, our simulations cover the entire range from weak to strong coupling. The analysis presented in this paper is primarily concerned with fundamental properties of superfluidity in strongly correlated spatially confined bosons in 2D and 3D traps. Examples include finite molecular bosonic clusters,\citep{khai,mezza,idowu,kwon,mebo} e.g.\ hydrogen or helium droplets, doublons in strongly correlated solids or ultracold lattice gases \citep{santos,rosch} as well as dipole-interacting trapped gases.\citep{dalf,mybo} In addition, there are numerous theoretical studies of the fundamental properties of the charged Bose gas (CBG) \citep{koscik,depalo,kim} and a variety of applications of the model to condensed matter physics. Indirect excitons \citep{lozovik,timo,sperlich} exhibit, despite their dipole-like interaction at long-range, a Coulomb-like repulsion at high density.\citep{boning} Furthermore, the $3D$ CBG of spatially bound electron pairs (small bipolarons) is discussed in the context of high $T_c$ superconductivity in cuprates.\citep{alexandrov,alexandrov2,lanzara,khasanov} Finally, applications also extend to macroscopic 3D bosonic Coulomb systems in the inner regions of neutron stars (proton pairing) \citep{peth,chamel,entrail} and the core of helium white dwarfs.\citep{white,white2} \section{\label{sec:level11}Model and simulation idea} We consider two- and three-dimensional systems of $N$ identical bosons in a harmonic confinement potential with frequency $\Omega$ which are described by the dimensionless Hamiltonian \begin{eqnarray} \label{hamiltonian}\hat{H} = -\frac{1}{2}\sum_{k=1}^N \nabla_k^2 + \frac{1}{2}\sum_{k=1}^N \mathbf{r}_k^2 + \frac{1}{2} \sum_{k\ne l}^N \frac{\lambda}{|\mathbf{r}_k - \mathbf{r}_l|} \quad , \end{eqnarray} and oscillator units (i.e., characteristic length $l_0=\sqrt{\hbar/m\Omega}$ and energy scale $E_0=\hbar\Omega$) are used throughout this work. The selected Coulomb repulsion in Eq.\ (\ref{hamiltonian}) serves as a test-case for a broad class of isotropic long-range interactions and the coupling constant $\lambda = q^2/(\hbar \Omega l_0)$ (with $q$ being the charge) can be controlled experimentally by the variation of the trap frequency. To simulate the system of interest at a particular inverse temperature $\beta=E_0/k_BT$ in thermodynamic equilibrium, we employ a realization of the widely used worm algorithm path integral Monte-Carlo method.\citep{wa} The latter delivers quasi-exact results by performing a Trotter decomposition of the density operator $\hat{\rho}=\textnormal{exp}(-\beta\hat{H})$ and each particle is represented by a path of $P$ positions (often denoted as ``beads'' or ``time slices'') in the imaginary time. To achieve sufficiently well converged results, we typically use $P=80\dots410$ while performing $N_\textnormal{MC}\sim 10^7$ independent measurements. \section{\label{sec:level111}Results} \subsection{\label{sec:level2}Superfluid crossover in $2D$ and $3D$} A convenient way to define superfluidity for finite sized, trapped systems is the consideration of the change of the moment of inertia due to quantum effects, also denoted as \textit{Hess-Fairbank}-effect. Within the Landau two-fluid model the total density $n$ is decomposed into a normal and a superfluid component, $n=n_\textnormal{n} + n_\textnormal{sf}$. Of particular interest is the superfluid fraction $\gamma_\textnormal{sf}$, i.e., the fraction of particles that do not participate in a rotation of the system. This quantity can be expressed in terms of the moments of inertia and, within the path integral picture, one can define an estimator as \citep{sf} \begin{eqnarray} \label{sf}\gamma_\textnormal{sf} = \frac{n_\textnormal{sf}}{n} = \frac{I_\textnormal{cl} - I}{I_\textnormal{cl}} = \frac{4m^2\braket{A_z^2}}{\beta\hbar^2I_\textnormal{cl}} \quad , \end{eqnarray} with the total and classical moment of inertia $I$ and $I_\textnormal{cl}$, respectively, and the area $\mathbf{A}$ which is enclosed by the particle trajectories \begin{eqnarray} \label{area} \mathbf{A} = \frac{1}{2}\sum_{k=1}^N\sum_{i=1}^{P} \left(\mathbf{r}_{k,i} \times \mathbf{r}_{k,i+1}\right) \quad . \end{eqnarray} In Eq.\ (\ref{sf}) the system is assumed to be set into rotation around the $z$-axis and, hence, only the area in the $x$-$y$-plane $A_z$ is relevant. \begin{figure} \centering \includegraphics[width=0.48\textwidth]{gamma1.pdf}\\ \vspace{-0.15cm} \includegraphics[width=0.48\textwidth]{gamma2.pdf}\\ \vspace{-1.cm} \includegraphics[width=0.48\textwidth]{gamma3.pdf} \caption{\textit{\textbf{Superfluid crossover in $2D$ and $3D$:}} The superfluid fraction $\gamma_\textnormal{sf}$ is plotted over the inverse temperature $\beta$ for $N=150$ particles in $2D$ (red triangles) and $3D$ (blue squares) for the coupling constant $\lambda=10$ (A), $\lambda=3$ (B) and $\lambda=1$ (C).} \label{picture} \end{figure} The results for the superfluid fraction, Eq.\ (\ref{sf}), are shown in Fig.\ \ref{picture}, where $\gamma_\textnormal{sf}$ is plotted versus the inverse temperature $\beta$ for $N=150$ particles in $2D$ (red triangles) and $3D$ (blue squares) and for the coupling parameters $\lambda=10$ (A), $\lambda=3$ (B) and $\lambda=1$ (C). All curves exhibit the expected increase of superfluidity with increasing $\beta$ and saturate to unity, i.e., a completely superfluid system. In addition, the crossover \cite{phase_trans_finite} is shifted to lower temperature with increasing particle interaction. This is a direct consequence of the increased inter-particle distance, which reduces the probability of the occurence of exchange cycles. However, it is interesting to note that, for all $\lambda$, the onset of the crossover occurs at lower temperature for the $3D$ system. This is a non-trivial observation and could be caused by different effects: (i) The availability of the additional dimension leads to a reduced degeneracy. This decreases exchange effects and hence, explains the observed shift of the onset superfluidity. (ii) The three-dimensional nature of the particle exchange could lead to a reduction of the projection of the area-vector $\mathbf{A}$ from Eq.\ (\ref{area}) onto a particular plane. In this case, one would observe the onset of the crossover at an increased degeneracy compared to the $2D$ system with the same parameters. To decide which (if any) of the two explanations is most likely we analyze the probability $P(L)$ for a single bead to be involved in an exchange cycle which consists of $L$ particles and, in addition, a degeneracy parameter $\chi$ [cf. Eq.\ (\ref{chi})]. For (i) to be correct, we expect similar values of $\chi$ and $P(L)$, for the same amount of superfluidity in the system. For (ii), on the other hand, we would expect a significantly increased $\chi$ and higher probabilities $P(L)$ for $L>1$, for equal $\gamma_\textnormal{sf}$ in $3D$. In Fig.\ \ref{picture_prob} the product $P(L)L$ (with $\sum_L P(L)L = 1$) is plotted versus $L$ again for $N=150$ particles and $\lambda=3$ for both $2D$ (red) and $3D$ (blue). The top image corresponds to an equal inverse temperature $\beta=2$, i.e., two systems with a non-zero and non-unity superfluid fraction $\gamma_\textnormal{sf}(2D)\approx0.92$ and $\gamma_\textnormal{sf}(3D)\approx0.34$. Both curves exhibit a similar decay with increasing $L$ and a sharp bend for very large exchange cycles, which occurs for smaller particle numbers in $3D$. In the $3D$ system both single and two particle trajectories are more probable than in $2D$ and the two curves intersect at $L=3$. All larger exchange cycles are significantly more probable in $2D$. The bottom image of Fig.\ \ref{picture_prob} shows the same information but for a fixed superfluid fraction $\gamma_\textnormal{sf}\approx0.6$. Here, the probability to be involved in a particular exchange cycle is nearly equal for both dimensionalities and the two curves intersect several times. In addition, the sharp bend for large $L$ occurs at the same position. This is a first hint towards explanation (i) because the same amount of superfluidity requires comparable realization rates of exchange cycles while, at the same inverse temperature, exchange is suppressed in $3D$ which explains the later onset of superfluidity compared to $2D$. For completeness, we note that, both $\lambda=1$ and $\lambda=10$, exhibit the same behavior as shown for $\lambda=3$ in Fig.\ \ref{picture_prob}. \vhalffig{perm.pdf}{Probability distribution of exchange cycles in $2D$ and $3D$}{The probability for a single bead to be involved in an exchange cycle of length $L$, $P(L)L$, is plotted over the former for $N=150$ particles and $\lambda=3$. The top image corresponds to $\beta=2$ with $\gamma_\textnormal{sf}(2D)\approx0.92$ and $\gamma_\textnormal{sf}(3D)\approx0.34$ and for the bottom image the inverse temperature has been adjusted that $\gamma_\textnormal{sf}\approx0.6$ in both dimensionalities, i.e., $\beta(2D)=1.35$ and $\beta(3D)=2.27$.}{picture_prob}{0.48} Next, we define the degeneracy parameter as \begin{eqnarray} \label{chi} \chi = \overline{n}\lambda_\beta^d \quad , \end{eqnarray} with the dimensionality $d$, the thermal de-Broglie wavelength $\lambda_\beta=\hbar \sqrt{2\pi\beta/m}$ and the mean density $\overline{n}$ (in $2D$ and $3D$, we compute the radial density $n$ by averaging over angular and spherical segments, respectively, with a fixed distance from the trap center), which has been averaged over the radial extension of the particular system. Thus, Eq.\ (\ref{chi}) provides a measure for the average number of particles within the approximate extension of a single-particle wavefunction. It should be noted that the strong inhomogeneity of the density of correlated trapped quantum particles, cf. Fig.\ \ref{picture3}, makes the definition in Eq.\ (\ref{chi}) arbitrary to some degree. However, we found that using $\overline{n}$ gives reliable results even for different shell structures. \twohalffig{all.pdf}{Degeneracy parameter}{The degeneracy parameter $\chi$ is plotted over the inverse temperature $\beta$ for the systems from Fig.\ \ref{picture}. The critical value $\chi_\textnormal{crit}$ (orange triangles) marks the superfluid crossover, i.e., $\gamma_\textnormal{sf}=0.5$. Image A covers the entire inverse temperature range of our simulations and image B shows a magnified segment around the superfluid crossover.}{picture2}{0.48}{zoom.pdf} The results for Eq.\ (\ref{chi}) are shown in Fig.\ \ref{picture2} (A) where $\chi$ is plotted versus $\beta$ for the systems from Fig.\ \ref{picture}. The degeneracy increases with $\beta$ for all simulated systems, as expected. However, a comparison between $2D$ and $3D$ systems with otherwise equal parameters reveals that, at high $T$, the three-dimensional system exhibits a smaller $\chi$, until the two curves intersect. This is an immediate consequence of the definition \ (\ref{chi}), where the de-Broglie wavelength enters with the power of the dimensionality $d$. This implies that, for a constant mean density $\overline{n}$, which becomes valid with increasing $\beta$, it is $\chi(d=2)\propto\beta$ and $\chi(d=3)\propto\beta^{3/2}$, i.e., a linear and polynomial behavior with respect to the inverse temperature in $2D$ and $3D$, respectively. The connection between Fig.\ \ref{picture2} and the superfluid crossover from Fig.\ \ref{picture} is given by the orange triangles, which mark the critical degeneracy $\chi_\textnormal{crit}$, i.e., the degeneracy parameter at the inverse temperature for which $\gamma_\textnormal{sf}=0.5$. Our simulations have revealed that, despite the onset of superfluidity at significantly lower temperature in $3D$, $\chi_\textnormal{crit}$ takes similar values for both dimensionalities. In fact, the $2D$ systems from Fig.\ \ref{picture2} even exhibit a slightly higher $\chi$ for the critical superfluid fraction than their $3D$ counterparts, cf.\ Fig.\ \ref{picture2} (B), where a magnified segment around the crossover is shown. However, this rather peculiar feature should not be over-interpretated due to the average character of the definition of the degeneracy parameter itself. Thus, we conclude that the behavior observed in Fig.\ \ref{picture2} rules out explanation (ii), and the different critical temperatures for the superfluid crossover in $2D$ and $3D$ appear to be a degeneracy effect. \subsection{Local superfluid density\label{seclsf}} Another interesting question is how the superfluidity is distributed across the system in the vicinity of the crossover. To investigate this topic we use a spatially resolved superfluid density estimator of Kwon \textit{et al.},\citep{kwon} that is consistent with the two-fluid interpretation: \begin{eqnarray} \label{nsf} n_\textnormal{sf}(\mathbf{r}) = \frac{4m^2}{\beta\hbar^2 I_\textnormal{cl}}\braket{A_zA_{z,\textnormal{loc}}(\mathbf{r})} \quad . \end{eqnarray} Here, $A_{z,\textnormal{loc}}(\mathbf{r})$ denotes a local contribution to the total area enclosed by the particle paths. In $2D$, the system is assumed to rotate around the axis perpendicular to the trap and the quantity from Eq.\ (\ref{nsf}) can be radially averaged without loosing information. For a $3D$ trap, however, the axis of rotation causes a break of the spherical symmetry and the definition of a meaningful average is less obvious. In this work, we average $n_\textnormal{sf}$ over spherical segments with the distance $r$ from the center of the trap. This gives a quantity that approaches the total radial density $n$ for a completely superfluid system, i.e., for $\gamma_\textnormal{sf}=1$. Nevertheless, one should keep in mind that, in general, particles with the same distance from the center $|{\bf r}|$ have different contributions to the classical moment of inertia, depending on the orientation of their radius vector ${\bf r}$ with respect to $\mathbf{e}_z$. \vhalffig{lsfraction.pdf}{Local superfluid fraction in $2D$ and $3D$}{The spatially resolved ratio of the superfluid and total density $\gamma_\textnormal{sf}(r)=n_\textnormal{sf}(r)/n(r)$ is plotted over the distance to the center of the trap $r$ for $N=150$ particles in $2D$ (top) and $3D$ (bottom) for a fixed superfluid fraction $\gamma_\textnormal{sf}\approx0.6$. The larger errors around the center of the trap are due to the worse statistics in small angular and spherical segments in $2D$ and $3D$, respectively.}{lsfraction}{0.48} \conffig{density.pdf}{Spatially resolved superfluidity}{The radial density $n$ (red triangles) and superfluid density $n_\textnormal{sf}$ (blue squares) are plotted with respect to the distance to the center of the trap $r$ for a fixed total superfluid fraction $\gamma_\textnormal{sf}\approx0.6$ for the systems from Figs.\ \ref{picture} and \ref{picture2}. The black curves correspond to a constant local SF ratio $\gamma_\textnormal{sf}(r)=0.6$.}{picture3}{0.90} Fig.\ \ref{lsfraction} shows the local ratio $\gamma_\textnormal{sf}(r)=n_\textnormal{sf}(r)/n(r)$ for $2D$ (top) and $3D$ (bottom) systems and an inverse temperature that has been choosen in a way that $\gamma_\textnormal{sf}\approx0.6$.\citep{adjustgamma} For both dimensionalities $\gamma_\textnormal{sf}(r)$ exhibits a decay with increasing distance from the center of the trap, $r$, which is most distinct for smaller coupling. This behavior can be understood by considering the normalization of the local superfluid density \begin{eqnarray} \int d\mathbf{r}\ n_\textnormal{sf}(\mathbf{r}) \mathbf{r}_\perp^2 = \gamma_\textnormal{sf} I_\textnormal{cl} \qquad . \label{nsfnorm} \end{eqnarray} Eq.\ (\ref{nsfnorm}) implies that $n_\textnormal{sf}$ integrates to the quantum correction to the moment of inertia, i.e., the missing contribution due to the particles in the superfluid phase. However, this also means that \begin{eqnarray} \frac{1}{N}\int d\mathbf{r}\ n_\textnormal{sf}(\mathbf{r}) \ne \gamma_\textnormal{sf} \quad , \end{eqnarray} because, for the global SF fraction, not only the ratio of superfluid and total particle numbers but, in addition, the position of a particular particle is relevant. According to Eq.~(\ref{nsfnorm}), particles near the center of the trap have a small contribution to the total moment of inertia and so their superfluidity only slightly influences $\gamma_\textnormal{sf}$. At the boundary, however, each particle has a significant contribution to $I$ and, thus, $n_\textnormal{sf}(r)$ in this region crucially influences the global SF fraction. With increasing $\lambda$, the interparticle repusion grows and the system extends radially (cf.\ Fig.\ \ref{picture3}), so this effect becomes even more important. For a fixed $\gamma_\textnormal{sf}$, as it is the case in Fig.\ \ref{lsfraction}, the spatially resolved SF fraction in the outermost shell must approach the global value, with increasing coupling strength. A similar effect has recently been observed by Kulchytskyy \textit{et al.} \citep{kulch} from PIMC simulation of $^4$He in a cylindrical nanopore where the local superfluid response has been obtained, both, with the area formula, Eq.\ (\ref{nsf}), and a winding number approach \citep{khai2} for a rotation around and the flow along the pore axis, respectively. Despite a very similar spatial distribution for the two estimators, there occurs a significant difference in the global SF fraction due to the different normalizations. The comparison between the $2D$ and the $3D$ results in Fig.\ \ref{lsfraction} reveals that, for $d=3$ and equal $\lambda$, the local superfluid fraction in the outer region exceeds its two dimensional counterpart. In fact, the uppermost (red) curve ($\lambda=10$) in the bottom image appears to be above the global value \citep{adjustgamma} almost everywhere, which seems to violate Eq.\ (\ref{nsfnorm}). However, this is a consequence of the symmetry break due to the external rotation around $\mathbf{e}_z$ and the applied averaging over spherical segments in $3D$. For this reason, all particles in the outermost shell have different contributions to the moment of inertia. The expected decrease of $\gamma_\textnormal{sf}(r)$ for the largest $\mathbf{r}_\perp$ is therefore masked in Fig.\ \ref{lsfraction} by contributions near $\mathbf{e}_z$ from the same shell. The remaining open question is why $\gamma_\textnormal{sf}$ is distributed inhomogeneously in the first place, in particular, why there is more superfluidity around the center of the trap than in the outer regions. In Fig.\ \ref{picture3}, both the total (red triangles) and superfluid density (blue squares) are plotted with respect to $r$ for the systems from Fig.\ \ref{lsfraction}. The top row shows density profiles for rather strong coupling, $\lambda=10$, for both the two- (left) and three-dimensional (right) system. In $2D$, the total density $n$ exhibits a relatively smooth decay with increasing $r$ until, at the boundary of the system, two shell-like oscillations appear. In $3D$, $n$ stays almost constant around the center and a very pronounced shell appears at the boundary. The center and bottom rows show the same information for medium ($\lambda=3$) and weak ($\lambda=1$) coupling, respectively. One clearly sees that the shell-like features disappear for both dimensionalities when the coupling is reduced. A general feature for all couplings is that, $n$ always decays in 2D whereas, in 3D, it is nearly constant in the center. Also the peak(s) at the boundary are always significantly stronger in three dimensions. The local superfluid density, on the other hand, exhibits a rather surprising behavior. For $d=2$ and strong coupling, $n_\textnormal{sf}$ equals $n$, around the center of the trap, whereas it is significantly suppressed for larger $r$, in particular within the outermost shell. The $3D$ counterpart exhibits a similar behavior. The two innermost, weakly pronounced shells are nearly completely superfluid, whereas the difference between $n$ and $n_\textnormal{sf}$ is the largest in the outermost shell. This non-uniformity of the superfluidity can be made more transparent by comparing to the case of a constant local superfluid fraction $\gamma_\textnormal{sf}(r)=0.6$ which is depicted by the black curves in Fig.\ \ref{picture3}. Clearly, the true superfluidity is always above (below) the black curve in the center (at the edge). Let us discuss the origin of this spatial variation of the superfluid density. The first reason that comes to mind is the spatial variation of the local degeneracy parameter which is proportional to the local total density, $\chi(r)\sim n(r)$. However, the numerical results for $n_\textnormal{sf}$ clearly deviate from this suggestion. In particular, for $\lambda=1, 3$ the superfluid density does not follow the local degeneracy (total density). Only at $\lambda=10$ when shells are formed, $n_\textnormal{sf}$ increases around the outermost shell. However, the absolute value of the superfluid density does not reach a maximum despite the maximum of the degeneracy parameter. From this we conclude that the spatial distribution of superfluidity is not governed by the spatial variation of degeneracy but by the local strength of correlations. In fact, it is well known that spatial order leads to a reduction of superfluidity. In Coulomb systems, particles mainly experience interaction with neighbors with smaller $r$, whereas most of the pair interactions with particles from the outer parts of the system cancel. This is a pure screening effect (Faraday cage effect), and cancellation would be complete in mean field approximation. Hence, the investigated Coulomb clusters clearly exhibit the strongest order around the boundary and are less correlated around the center of the trap. This is also manifest in the shell formation at the outer boundary (red curves in Fig.\ \ref{picture3}) which is observed in classical Coulomb clusters as well.\citep{hanno} Another mechanism that suppresses superfluidity at the cluster boundary is a geometrical effect: within the outermost shell there are fewer particles available for the formation of exchange cycles due to the lack of neighbors in radial direction. The medium and weakly coupled system in $2D$ exhibit the same trend as for $\lambda=10$. In $3D$, the largest deviation between $n$ and $n_\textnormal{sf}$ appears around the outermost shell as well, whereas the superfluid density remains nearly constant over the rest of the system. This is in contrast to $2D$ and arises from the different behaviors of the total density $n(r)$. At the same time, the spatially resolved SF fraction (cf.\ Fig.\ \ref{lsfraction}) exhibits similar trends in $2D$ and $3D$. \section{Discussion} In summary, we have presented ab-initio results for the superfluid properties of Coulomb interacting bosons in both $2D$ and $3D$ harmonic traps. It was revealed that the availability of an additional dimension causes a decrease of the critical temperature of the superfluid crossover. To explain this non-trivial feature, we have analyzed the probability distribution of exchange cycles and the degeneracy parameter (\ref{chi}), and it was revealed that the critical superfluid fraction, $\gamma_\textnormal{sf}=0.5$, occurs for similar degeneracy for both dimensionalities, despite the significantly higher $\beta$ in $3D$. In addition, we have investigated the spatial distribution of superfluidity across the system by using a local superfluid density estimator. In both $2D$ and $3D$, the largest difference between the total and superfluid density $n$ and $n_\textnormal{sf}$, occurs near the boundary of the system. This is a direct consequence of the increased order at large distances from the center of the trap. The onset of superfluidity at comparatively lower temperature in $3D$ systems is expected to be a general feature and not specific for Coulomb interaction. For the investigated particle number $N=150$, superfluidity requires a collective response of the entire system, in particular the frequent realization of large exchange cycles in the path integral picture. The probability for the latter crucially depends on the degeneracy. The reported $\beta$-dependence of the degeneracy parameter $\chi$ is valid for other long-range interactions as well, because the coupling only enters in the average density $\overline{n}$ which remains constant with decreasing $T$. The observed non-uniform distribution of superfluidity in the vicinity of the crossover, on the other hand, depends on several system properties. For small $N$ and strong coupling, the exact particle number plays an important role and, in case of strong hexagonal order at the center of the trap, the superfluidity is essentially located at the boundary.\citep{fili} The almost complete screening of the interaction of a given particle with particles located at a larger distance from the center is a Coulomb specific effect. In the case of other pair interactions, a force from the outer regions may exist, which leads to different density profiles and spatial distributions of correlation effects.\citep{hanno} In addition, the explicit choice of the confinement potential also influences these quantities. The application of e.g. a quartic trap is expected to enhance the behavior observed in Sec.\ (\ref{seclsf}), whereas a weaker than harmonic confinement is expected to reduce these trends. While we expect that the observed trends are typical for Bose systems in confinement potentials, we mention that, for unconfined finite systems with attractive interaction, different behaviors have been predicted. Khairallah \textit{et al.} reported that, for mesoscopic parahydrogen clusters, superfluidity could be realized by loosely bound surface molecules,\citep{khai} although this is in disagreement with results by Mezzacapo and Boninsegni.\citep{mezza} \section*{\small Acknowledgements} This work is supported by the Deutsche Forschungsgemeinschaft via SFB-TR24 and project FI 1252/2 and by grant SMP006 for CPU time at the HLRN.
\section{Introduction\protect\\} Fe-Cr-Ni alloys are one of the most studied ternary alloy systems. Their significance stems from the fact that they form the basis for many types of austenitic, ferritic and martensitic steels. Ternary Fe-Cr-Ni and binary Fe-Cr, Fe-Ni and Ni-Cr alloys exhibit diverse magnetic, thermodynamic and mechanical properties, which make them suitable for a variety of applications. This alloy family includes several outstanding examples, like Invar\cite{Guillaume1897} and Permalloy\cite{Arnold1923}. Fe-Cr-Ni based steels, including austenitic 304 and 316 steels, are widely used as structural materials for light water and fast breeder fission reactors\cite{Klueh,Toyama2012}. Inconel alloys X-750 and 718 are used in reactor core components\cite{Rowcliffe2009}. Fe-Cr-based steels F82H and Eurofer are among candidate structural materials for tritium breeding blankets of fusion reactors \cite{Stork2014}. Since the stability of materials in extreme conditions is affected by many factors, extensive and accurate knowledge of how materials respond to temperature and irradiation over extended periods of time is required. The selection of optimal alloy compositions is therefore one of the objectives of fission and fusion materials research. For example, there is a perception that bcc alloys like V-Cr-Ti alloys or ferritic steels exhibit better resistance to radiation swelling in comparison with fcc alloys \cite{Boutard2008}. However, it has been shown by Satoh \textit{et al.} \cite{Satoh2007} that in the fcc Fe$_{55}$Cr$_{15}$Ni$_{30}$ alloy irradiated up to 6 dpa swelling is also significantly reduced when temperature is above 350$^{\circ}$C. Because of the broad range of applications of Fe-Cr-Ni alloys, their phase diagram has been extensively assessed from the thermodynamic perspective. Microstructure of Fe-Cr-Ni steels is well described by the Schaeffler diagram\cite{Ferry2006}. The phase composition of steels can be controlled by varying Cr and Ni content, since chromium is a ferrite (bcc phase) stabilizer and nickel is an austenite (fcc phase) stabilizer. A thermodynamic model for Fe-Cr-Ni alloys employing CALPHAD method has been developed using interpolation of elevated temperature experimental data\cite{Rees1949,Hattersley1966,Cook1952}. Due to the relatively slow kinetics of relaxation towards equilibrium at low temperatures, the amount of experimental information about the low temperature part of the phase diagram is limited. This information can instead be derived from {\it ab initio} DFT simulations\cite{Koermann2014}, as was recently demonstrated for binary Fe-Ni alloys in Ref. \onlinecite{Cacciamani2010}. A recent revision of the Fe-Cr-Ni CALPHAD phase diagram is given in Ref. \onlinecite{Franke2011}, where both magnetic and chemical ordering temperatures of binary Fe-Ni alloys were extrapolated to ternary alloys. There have been only a few DFT studies of Fe-Cr-Ni ternary alloys. Properties of the alloys in the dilute Cr and Ni limit were analyzed in Refs. \onlinecite{Klaver2012,Hepburn2013}. The Coherent Potential Approximation (CPA) was used by the authors of Refs. \onlinecite{Vitos2002,Vitos2006,Delczeg2012}. Recently\cite{Piochaud2014}, Special Quasi-random Structures (SQS) \cite{Zunger1990} were used for investigating point defects in fcc Fe$_{70}$Cr$_{20}$Ni$_{10}$ alloys. In all these studies, Fe-Cr-Ni alloys were assumed to be fully chemically disordered. This assumption is not realistic, since there is direct experimental evidence showing that many Fe-Cr-Ni alloys exhibit short-range order\cite{Dimitrov1986,Cenedese1984,Menshikov1997}. Whilst chemical SRO is naturally expected for ternary alloy compositions close to the known binary intermetallic phases like FeNi$_3$, FeNi and CrNi$_2$, SRO in FeNi$_3$ alloyed with Cr is found to decrease rapidly as a function of Cr content \cite{Marwick1987}. Unexpectedly, a significant degree of chemical order is observed in alloys with compositions very different from that of binary intermetallic phases, for example in Fe$_{56}$Cr$_{21}$Ni$_{23}$ \cite{Cenedese1984}, Fe$_{64}$Cr$_{16}$Ni$_{20}$, Fe$_{59}$Cr$_{16}$Ni$_{25}$ \cite{Dimitrov1986} and Fe$_{34}$Ni$_{46}$Cr$_{20}$ \cite{Menshikov1997}. Chemical order in alloys, and various properties of ordered alloys, can be analyzed using a combination of first-principles calculations and statistical mechanics simulations based on a generalization of the Ising alloy model. In the CE model, the energy of an alloy is represented by a series in cluster functions, where the resulting expression for the energy has the form of a generalized Ising Hamiltonian containing several coupling parameters known as Effective Cluster Interactions (ECIs)\cite{Sanchez1984}. Various methods have been developed to compute ECIs from first principles. The most often used is the Structure Inversion Method (SIM), based on the Connolly-Williams approximation \cite{Connolly1983}, and the coherent potential approximation used in combination with the Generalized Perturbation Method (CPA-GPM). In the CPA-GPM scheme, a random alloy is constructed by considering average occupancies of lattice sites by atoms of alloy components, where coupling parameters are computed using a perturbation approach \cite{Ruban2008}. In SIM, energies of ordered structures are computed using DFT, and then ECIs are obtained through least-squares fitting. Both techniques have been successfully applied to binary alloy sub-systems of Fe-Cr-Ni \cite{Klaver2006,Nguyen-Manh2007,Nguyen-Manh2012,Lavrentiev2007,Ruban2008,Barabash2009,Ekholm2010,Rahaman2014}. However, ternary Fe-Cr-Ni alloys have not received attention. In this study we use SIM, since the accuracy of ECIs is primarily controlled by the approximations involved in {\it ab initio} calculations of energies of input structures, and by the cross-validation error between DFT and CE. The last but not least critical issue to consider here is the broad variety of magnetic configurations characterizing fcc and bcc Fe-Cr-Ni alloys. For example, fcc Fe$_{80-x}$Ni$_x$Cr$_{20}$ alloys ($10<x<30$) exhibit ferromagnetic, anti-ferromagnetic, or spin-glass type magnetic order, or a mixture of all of them \cite{Majumdar1984}. To find the most stable atomic structures needed for parameterizing the CE model, many magnetic configurations were computed and their energies compared. Variation of magnetic properties as functions of alloy composition was investigated, including the occurrence of magneto-volume effects in Fe-Cr-Ni alloys. Effective cluster interaction parameters, obtained by mapping DFT energies of stable collinear magnetic configurations to CE, are used in quasi-canonical MC simulations. Here we investigate the phase stability and chemical order of fcc and bcc Fe-Cr-Ni alloys at finite temperatures and generate representative alloy structures for future DFT analysis of radiation defects in alloys. We also analyze magnetic properties of Fe-Cr-Ni alloys at low and high temperatures using MCE-based Monte Carlo simulations. The paper is structured as follows. In Section II, we describe the CE formalism for multi-component alloys, focusing on the ternary alloy systems, and derive formulae for short-range order parameters expressed in terms of cluster functions. In Section III we analyze the phase stability and magnetic properties of alloy structures predicted by DFT at 0 K. Finite temperature phase stability and chemical order are investigated using quasi-canonical MC simulations in Section IV. Finite-temperature magnetic properties are explored by MCE simulations in Sections V. Conclusions are given in Section VI. \section{Computational methodology\protect\\} \subsection{Cluster expansion formalism for ternary alloys} The stability of ternary alloy phases can be investigated using a combination of quantum-mechanical DFT calculations and lattice statistical mechanics simulations. The enthalpy of mixing of an alloy, which can be evaluated using DFT, is defined as \begin{eqnarray} \Delta H^{lat}_{DFT} (\vec{\sigma})&=&E^{lat}_{tot}(A_{c_B}B_{c_B}C_{c_C},\vec{\sigma})-c_AE^{lat}_{tot}(A) \nonumber \\ &-&c_BE^{lat}_{tot}(B)-c_CE^{lat}_{tot}(C), \label{eq:Mixing_DFT} \end{eqnarray} where $c_A$, $c_B$ and $c_C$ are the average concentrations of alloy components A, B and C. $E^{lat}_{tot}$ are the total energies of relevant structures defined assuming a certain crystal lattice. Superscript $lat$ denotes the chosen lattice type: face-centred cubic (fcc) or body-centred cubic (bcc). An atomic alloy configuration is specified by a vector of configurational variables $\vec{\sigma}$. In cluster expansion, the configurational enthalpy of mixing of a ternary alloy is defined as \cite{Walle2009} \begin{equation} \Delta H_{CE}(\vec{\sigma}) = \sum_{\omega}m_{\omega}J_{\omega}\left\langle \Gamma_{\omega'}(\vec{\sigma})\right\rangle_\omega , \label{eq:CE_1} \end{equation} where summation is performed over all the clusters $\omega$ that are distinct under group symmetry operations of the underlying lattice, $m^{lat}_\omega$ are multiplicity factors indicating the number of clusters equivalent to $\omega$ by symmetry (divided by the number of lattice sites), $\left\langle \Gamma_{\omega'}(\vec{\sigma})\right\rangle$ are the cluster functions defined as products of \textit{functions} of occupation variables on a specific cluster $\omega$ averaged over all the clusters $\omega'$ that are equivalent by symmetry to cluster $\omega$. $J_{\omega}$ are the concentration-independent Effective Cluster Interaction (ECI) parameters, derived from a set of \textit{ab-initio} calculations using the structure inversion method \cite{Connolly1983}. A cluster $\omega$ is defined by its size (number of lattice points) $|\omega|$, and the relative positions of points. Coordinates of points in each cluster considered here for fcc and bcc lattices are listed in Table \ref{tab:ECI_def_ternary}. For clarity, each cluster $\omega$ is described by two parameters $(|\omega|,n)$, where $|\omega|$ is the cluster size and $n$ is a label, defined in Table \ref{tab:ECI_def_ternary}. In binary alloys, lattice site occupation variables are usually defined as $\sigma_i=\pm1$, where $\sigma$ indicates whether site \textit{i} is occupied by an atom of type A ($\sigma_i=+1$) or B ($\sigma_i=-1$). In this case the cluster function is defined as a product of occupation variables over all the sites included in cluster $\omega$ \begin{equation} \Gamma_{\omega,n}(\vec{\sigma}) = \sigma_1\sigma_2\ldots\sigma_{|\omega|}. \label{eq:CE_2} \end{equation} In a $K$-component system, a cluster function is not a simple product of occupation variables. Instead, it is defined as a product of orthogonal point functions $\gamma_{j_i,K}(\sigma_i)$, \begin{equation} \Gamma_{\omega,n}^{(s)}(\vec{\sigma}) = \gamma_{j_1,K}(\sigma_1)\gamma_{j_2,K}(\sigma_2)\ldots\gamma_{j_{|\omega|},K}(\sigma_{|\omega|}), \label{eq:CE_3} \end{equation} where sequence $(s) =(j_{1} j_{2} \ldots\ j_{|\omega|})$ is the {\it decoration} \cite{Sandberg2007} of cluster by point functions. All the decorations of clusters, which are not symmetry-equivalent for fcc and/or bcc ternary alloys, are given in Table \ref{tab:ECI_def_ternary} together with their multiplicities $m_{|\omega|,n}^{(s)}$ and effective cluster interactions $J_{|\omega|,n}^{(s)}$. The number of possible decorations of clusters by non-zero point functions is a permutation with repetitions, $\left.(K-1)^{|\omega|}\right.$. Effective cluster interactions for those clusters are given in Table \ref{tab:ECI_def_ternary} only once, together with the corresponding multiplicity factor $m_{|\omega|,n}$. In ternary alloys, occupation variables and point functions can be defined in various ways. For example, in Ref. \onlinecite{Sanchez1984,Wolverton1994} occupation variables are defined as $\sigma_i=-1,0,+1$ and point functions as: $\gamma_{0,3}=1$ (for the zero cluster), $\gamma_{1,3}(\sigma_i)=\sqrt{\frac{3}{2}}\sigma_i$, and $\gamma_{2,3}(\sigma_i)=\sqrt{2}(1-\frac{3}{2}\sigma_i^2)$. We define occupation variables and point functions following Ref. \onlinecite{Walle2009}. This allows us to apply the same formulae as for a $K$-component system \begin{equation} \gamma_{j,K}\left(\sigma_i\right)=\begin{cases} 1 & \textrm{ if }j=0\textrm{ }, \\ -\cos\left(2\pi\lceil\frac{j}{2}\rceil\frac{\sigma_i}{K}\right) & \textrm{ if }j>0\textrm{ and odd}, \\ -\sin\left(2\pi\lceil\frac{j}{2}\rceil\frac{\sigma_i}{K}\right) & \textrm{ if }j>0\textrm{ and even}, \end{cases} \label{eq:point_functions} \end{equation} where $\sigma_i = 0,1,2,\ldots,\left(K-1\right)$, $j$ is the index of point functions ($j=0,1,2,\ldots,(K-1)$), and where $\lceil \frac{j}{2} \rceil$ denotes an operation where we take the integer plus one value of a non-integer number, for example $\lceil 2.5 \rceil=3$. In ternary alloys, index $K$ equals 3. In what follows we will drop it to simplify notations. Occupation variables are now defined as $\sigma=0,1,2$, referring to the constituent components of the alloy \textit{A}, \textit{B} and \textit{C}, which here correspond to Fe, Cr, and Ni, respectively. The enthalpy of mixing (Eq. \ref{eq:CE_1}) of a ternary alloy on a lattice can now be written as \begin{widetext} \begin{eqnarray} \Delta H_{CE}(\vec{\sigma}) &=& \sum_{|\omega|,n,s}m_{|\omega|,n}^{(s)}J_{|\omega|,n}^{(s)}\left\langle \Gamma_{|\omega'|,n'}^{(s')}(\vec{\sigma})\right\rangle_{|\omega|,n,s} \nonumber \\ &=& J_{1,1}^{(0)}\left\langle \Gamma_{1,1}^{(0)}\right\rangle+J_{1,1}^{(1)}\left\langle \Gamma_{1,1}^{(1)}\right\rangle + J_{1,1}^{(2)}\left\langle \Gamma_{1,1}^{(2)}\right\rangle+\sum_{n=1}^{pairs} \left(m_{2,n}^{(11)}J_{2,n}^{(11)}\left\langle \Gamma_{2,n}^{(11)}\right\rangle + m_{2,n}^{(12)}J_{2,n}^{(12)}\left\langle \Gamma_{2,n}^{(12)}\right\rangle \right. \nonumber \\ &+& \left. m_{2,n}^{(22)}J_{2,n}^{(22)}\left\langle \Gamma_{2,n}^{(22)}\right\rangle\right) + \sum_{n=1}^{multibody} \ldots \label{eq:CE_expanded_1} \end{eqnarray} Expressions for fcc and bcc alloys differ because of their different multiplicity factors, $m_{|\omega|,n}^{(s)}$, given in Table \ref{tab:ECI_def_ternary}. \begin{center} \begin{longtable}{ccccccccc} \multicolumn{9}{c}{\parbox{12cm}{TABLE I. Size $|\omega|$, label $n$, decoration $(s)$, multiplicity $m_{|\omega|,n}^{(s)}$ and coordinates of points in the relevant clusters on fcc and bcc lattices. $J_{|\omega|,n}^{(s)}$ (in meV) are the effective cluster interaction parameters for fcc and bcc ternary Fe-Cr-Ni alloys. Index $(s)$ is the same as the sequence of points in the relevant cluster.}} \label{tab:ECI_def_ternary} \\ \hline \hline \multicolumn{3}{c}{ } & \multicolumn{3}{c}{fcc} & \multicolumn{3}{c}{bcc} \\ $|\omega|$ & $n$ & ($s$) & Coordinates & $m_{|\omega|,n}^{(s)}$ & $J_{|\omega|,n}^{(s)}$ & Coordinates & $m_{|\omega|,n}^{(s)}$ & $J_{|\omega|,n}^{(s)}$ \\ \hline \endfirsthead \multicolumn{9}{c}{ TABLE I. (\textit{Continued})} \\ \hline \hline \multicolumn{3}{c}{ } & \multicolumn{3}{c}{fcc} & \multicolumn{3}{c}{bcc} \\ $|\omega|$ & $n$ & ($s$) & Coordinates & $m_{|\omega|,n}^{(s)}$ & $J_{|\omega|,n}^{(s)}$ & Coordinates & $m_{|\omega|,n}^{(s)}$ & $J_{|\omega|,n}^{(s)}$ \\ \hline \endhead \hline \hline \endfoot \hline \hline \endlastfoot 1 & 1 & (0) & (0,0,0) & 1 & -77.281 & (0,0,0) & 1 & 132.945 \\ & & (1) & & 1 & -60.747 & & 1 & 47.929 \\ & & (2) & & 1 & 2.847 & & 1 & -168.929 \\ 2 & 1 & (1,1) & (0,0,0; $\frac{1}{2}$,$\frac{1}{2}$,0) & 6 & 4.329 & (0,0,0; $\frac{1}{2}$,$\frac{1}{2}$,$\frac{1}{2}$) & 4 & -54.656 \\* & & (1,2) & & 12 & -2.057 & & 8 & -4.140 \\* & & (2,2) & & 6 & -2.039 & & 4 & -64.784 \\ 2 & 2 & (1,1) & (0,0,0; 1,0,0) & 3 & -9.596 & (0,0,0; 1,0,0) & 3 & -19.159 \\* & & (1,2) & & 6 & 7.284 & & 6 & 7.332 \\* & & (2,2) & & 3 & -31.827 & & 3 & -19.253 \\* 2 & 3 & (1,1) & (0,0,0; 1,$\frac{1}{2}$,$\frac{1}{2}$) & 12 & 3.345 & (0,0,0; 1,0,1) & 6 & -1.547 \\* & & (1,2) & & 24 & -0.702 & & 12 & 11.871 \\* & & (2,2) & & 12 & 4.224 & & 6 & 8.392 \\ 2 & 4 & (1,1) & (0,0,0; 1,1,0) & 6 & -1.990 & (0,0,0; 1$\frac{1}{2}$,$\frac{1}{2}$,$\frac{1}{2}$) & 12 & 2.466 \\* & & (1,2) & & 12 & 1.192 & & 24 & 0.564 \\* & & (2,2) & & 6 & 6.662 & & 12 & -2.660 \\ 2 & 5 & (1,1) & (0,0,0; 1$\frac{1}{2}$,$\frac{1}{2}$,$\frac{1}{2}$) & 6 & -2.034 & (0,0,0; 1,1,1) & 4 & 1.602 \\* & & (1,2) & & 12 & 0.724 & & 8 & -1.368 \\* & & (2,2) & & 6 & 2.036 & & 4 & 3.031 \\ 3 & 1 & (1,1,1) & (0,0,0; $\frac{1}{2}$,0,$\frac{1}{2}$; & 8 & -9.015 & (1,0,0; $\frac{1}{2}$,$\frac{1}{2}$,$\frac{1}{2}$; & 12 & -6.961 \\* & & (2,1,1) & 0,$\frac{1}{2}$,$\frac{1}{2}$) & 24 & 3.847 & 0,0,0) & 24 & 8.827 \\* & & (1,2,1) & & & & & 12 & 1.620 \\* & & (2,2,1) & & 24 & -6.544 & & 24 & -1.954 \\* & & (2,1,2) & & & & & 12 & 22.895 \\* & & (2,2,2) & & 8 & 12.492 & & 12 & 2.934 \\ 3 & 2 & (1,1,1) & (1,0,0; $\frac{1}{2}$,-$\frac{1}{2}$,0; & 12 & -3.019 & ($\frac{1}{2}$,-$\frac{1}{2}$,-$\frac{1}{2}$; 0,0,0; & 12 & -6.255 \\* & & (2,1,1) & 0,0,0) & 24 & -0.470 & -$\frac{1}{2}$,-$\frac{1}{2}$,$\frac{1}{2}$) & 24 & 2.510 \\* & & (1,2,1) & & 12 & -1.778 & & 12 & -1.292 \\* & & (2,2,1) & & 24 & 5.371 & & 24 & 6.122 \\* & & (2,1,2) & & 12 & 6.310 & & 12 & 6.580 \\* & & (2,2,2) & & 12 & -0.126 & & 12 & 4.334 \\ 3 & 3 & (1,1,1) & ($\frac{1}{2}$,$\frac{1}{2}$,0; 0,0,0; & 24 & 0.821 & & & \\* & & (2,1,1) & -$\frac{1}{2}$,0,$\frac{1}{2}$) & 48 & -0.017 & & & \\* & & (1,2,1) & & 24 & 0.931 & & & \\* & & (2,2,1) & & 48 & 0.369 & & & \\* & & (2,1,2) & & 24 & 2.657 & & & \\* & & (2,2,2) & & 24 & -3.945 & & & \\ 4 & 1 & (1,1,1,1) & (0,0,0; $\frac{1}{2}$,$\frac{1}{2}$,0; & 2 & -12.978 & (1,0,0; $\frac{1}{2}$,-$\frac{1}{2}$,$\frac{1}{2}$; & 6 & -12.095 \\* & & (2,1,1,1) & $\frac{1}{2}$,0,$\frac{1}{2}$; 0,$\frac{1}{2}$,$\frac{1}{2}$) & 8 & -1.931 & $\frac{1}{2}$,$\frac{1}{2}$,$\frac{1}{2}$; 0,0,0) & 24 & -13.020 \\* & & (2,2,1,1) & & 12 & 4.987 & & 24 & 0.000 \\* & & (1,2,2,1) & & & & & 12 & 0.000 \\* & & (2,2,2,1) & & 8 & -1.140 & & 24 & 0.000 \\* & & (2,2,2,2) & & 2 & 0.824 & & 6 & 0.007 \\ 4 & 2 & (1,1,1,1) & (1,0,0; $\frac{1}{2}$,0,$\frac{1}{2}$; & 12 & -1.452 & & & \\* & & (2,1,1,1) & $\frac{1}{2}$,-$\frac{1}{2}$,0; 0,0,0) & 24 & 1.076 & & & \\* & & (1,2,1,1) & & 24 & -1.775 & & & \\* & & (2,2,1,1) & & 48 & 1.114 & & & \\* & & (1,2,2,1) & & 12 & -0.581 & & & \\* & & (2,2,2,1) & & 24 & -5.109 & & & \\* & & (2,1,1,2) & & 12 & 4.130 & & & \\* & & (2,2,1,2) & & 24 & 2.549 & & & \\* & & (2,2,2,2) & & 12 & 6.127 & & & \\ 5 & 1 & (1,1,1,1,1) & (1,0,0; $\frac{1}{2}$,0,-$\frac{1}{2}$; & 6 & 4.219 & (1,0,0; $\frac{1}{2}$,-$\frac{1}{2}$,$\frac{1}{2}$; & 12 & -6.356 \\* & & (2,1,1,1,1) & $\frac{1}{2}$,0,$\frac{1}{2}$; $\frac{1}{2}$,-$\frac{1}{2}$,0; & 24 & -1.263 & $\frac{1}{2}$,$\frac{1}{2}$,$\frac{1}{2}$; 0,0,0; & 24 & 7.696 \\* & & (1,2,1,1,1) & 0,0,0) & & & 0,0,1) & 24 & -15.998 \\* & & (2,2,1,1,1) & & 24 & 0.626 & & 48 & 15.385 \\* & & (1,2,2,1,1) & & 12 & 1.676 & & 12 & -20.341 \\* & & (2,2,2,1,1) & & 24 & -0.360 & & 24 & 14.846 \\* & & (1,1,1,2,1) & & 6 & -6.115 & & 12 & -5.003 \\* & & (2,1,1,2,1) & & 24 & -1.565 & & 24 & -3.067 \\* & & (1,2,1,2,1) & & & & & 24 & -3.221 \\* & & (2,2,1,2,1) & & 24 & 3.258 & & 48 & 1.070 \\* & & (1,2,2,2,1) & & 12 & 2.284 & & 12 & -3.473 \\* & & (2,2,2,2,1) & & 24 & -1.400 & & 24 & -1.255 \\* & & (2,1,1,1,2) & & & & & 12 & 11.683 \\* & & (2,2,1,1,2) & & & & & 24 & -1.192 \\* & & (2,2,2,1,2) & & 6 & 1.565 & & 12 & -2.460 \\* & & (2,1,1,2,2) & & & & & 12 & -6.855 \\* & & (2,2,1,2,2) & & & & & 24 & -7.050 \\* & & (2,2,2,2,2) & & 6 & -6.793 & & 12 & -5.397 \\* \end{longtable \end{center} \end{widetext} Configuration averages $\left\langle \Gamma_{|\omega|,n}^{(s)}(\vec{\sigma})\right\rangle$ in Eq. \ref{eq:CE_expanded_1} can be expressed in terms of point, pair and multi-body probabilities. An average point correlation function can be calculated using the equation \begin{equation} \left\langle \Gamma_{1,1}^{(s)}\right\rangle=\langle \gamma_j\rangle=\sum_{k=1}^3T_{jk}\times\langle p^{(k)}\rangle=\sum_{k=1}^3T_{jk}c_k , \label{eq:point_corr_function} \end{equation} where $k=0,1,2$, $p^{(k)}$ are the site-occupation operators counting the number of sites occupied by the same atom type \cite{Ducastelle1991}. Average values of site-occupation operators $\left\langle p^{(k)}\right\rangle=c_k$ are concentrations $c_A$, $c_B$ and $c_C$, and $T_{ij}$ are elements of the point probability matrix given, through Eq. \ref{eq:point_functions}, by \begin{equation} \left[\begin{array}{c} \left\langle \gamma_0 \right\rangle \\ \left\langle \gamma_1 \right\rangle \\ \left\langle \gamma_2 \right\rangle \\ \end{array} \right]= \left[ \begin{array}{ccc} 1 & 1 & 1 \\ -1 & \frac{1}{2} & \frac{1}{2} \\ 0 & -\frac{\sqrt{3}}{2} & \frac{\sqrt{3}}{2} \\ \end{array} \right] \left[\begin{array}{c} \left\langle p^{(0)} \right\rangle \\ \left\langle p^{(1)} \right\rangle \\ \left\langle p^{(2)} \right\rangle \\ \end{array} \right]. \label{eq:CE_M_Matrix} \end{equation} The three average point functions are therefore \begin{eqnarray} \left\langle \Gamma_{1,1}^{(0)}\right\rangle&=&\langle\gamma_0\rangle = \sum_ic_i\gamma_0(\sigma_i)= 1 \nonumber \\ \left\langle \Gamma_{1,1}^{(1)}\right\rangle&=&\langle\gamma_1\rangle =\sum_ic_i\gamma_1(\sigma_i)=\frac{1}{2}\left(-2c_A+c_B+c_C\right)=\frac{1}{2}\left(1-3c_A\right) \nonumber \\ \left\langle \Gamma_{1,1}^{(2)}\right\rangle&=&\langle\gamma_2\rangle = \sum_ic_i\gamma_2(\sigma_i)=\frac{\sqrt{3}}{2}\left(c_C-c_B\right) . \label{eq:Gamma_point_clust} \end{eqnarray} Similarly to Eq. \ref{eq:point_corr_function}, the average cluster functions for pairwise clusters ($n$-th nearest neighbours) are linear functions of the average pairwise probabilities. They are given by \begin{eqnarray} \left\langle \Gamma_{2,n}^{(ij)}\right\rangle&=&\langle \gamma_i,\gamma_j\rangle_{n}=\sum_{h=1}^3\sum_{k=1}^3T_{ih}T_{jk}\times\langle p^{(h)}p^{(k)}\rangle_{n} \nonumber \\ &=&\sum_{h=1}^3\sum_{k=1}^3\gamma_i(\sigma_h)\gamma_j(\sigma_k)y_n^{hk}, \label{eq:pair_cluster_funct} \end{eqnarray} where $T_{ih}$ and $T_{jk}$ are elements of the point probability matrix (Eq. \ref{eq:CE_M_Matrix}), and $y_n^{hk}$ is the temperature-dependent probability of finding atom $h$ near atom $k$ in the $n$-th nearest neighbour coordination shell, given by \cite{Ducastelle1991} \begin{equation} y_n^{hk}=\langle p^{(h)}p^{(k)}\rangle_{n}=\langle p^{(h)}\rangle\langle p^{(k)}\rangle\left(1-\alpha_n^{hk}\right)=c_hc_k\left(1-\alpha_n^{hk}\right). \label{eq:pair_probability} \end{equation} Here $\alpha_n^{hk}$ is the Warren-Cowley short-range parameter for atoms $h$ and $k$ in the $n$-th neighbour shell, defined as the deviation from entirely random distribution of atoms in the alloy. Average cluster functions for the three pairs of non-equivalent atoms are therefore \begin{eqnarray} \left\langle \Gamma_{2,n}^{(11)}\right\rangle&=&\langle\gamma_1,\gamma_1\rangle_{n}= \nonumber \\ &=&\frac{1}{4}\left(1+3y_n^{AA}-6y_n^{AB}-6y_n^{AC}\right) \nonumber \\ \left\langle \Gamma_{2,n}^{(12)}\right\rangle&=&\langle\gamma_1,\gamma_2\rangle_{n}= \nonumber \\ &=&\frac{\sqrt{3}}{4}\left(-y_n^{BB}+y_n^{CC}+2y_n^{AB}-2y_n^{AC}\right) \nonumber \\ \left\langle\Gamma_{2,n}^{(22)}\right\rangle&=&\langle\gamma_2,\gamma_2\rangle_{n}=\frac{3}{4}\left(y_n^{BB}+y_n^{CC}-2y_n^{BC}\right). \label{eq:Gamma_pairs} \end{eqnarray} Rewriting Eq. \ref{eq:CE_expanded_1} in terms of average point and pair functions given by Eqs. \ref{eq:Gamma_point_clust} and \ref{eq:Gamma_pairs}, we find that the configurational enthalpy of mixing for a ternary alloy can be expressed as a function of concentrations $c_i$ and average pair probabilities $y_n^{ij}$ via \begin{widetext} \begin{eqnarray} \Delta H_{CE}(\vec{\sigma}) &=& J_1^{(0)}+J_1^{(1)}\left(1-3c_A\right) + J_1^{(2)}\frac{\sqrt{3}}{2}\left(c_C-c_B\right) \nonumber \\ &+&\sum_{n}^{pairs} \left[\frac{1}{4}m_{2,n}^{(11)}J_{2,n}^{(11)}\left(1+3y_n^{AA}-6y_n^{AB}-6y_n^{AC}\right) + \frac{\sqrt{3}}{4}m_{2,n}^{(12)}J_{2,n}^{(12)}\left(-y_n^{BB}+y_n^{CC}+2y_n^{AB}-2y_n^{AC}\right) \right. \nonumber \\ &+& \left.\frac{3}{4}m_{2,n}^{(22)}J_{2,n}^{(22)}\left(y_n^{BB}+y_n^{CC}-2y_n^{BC}\right) \right] + \sum_{n}^{multibody} \ldots \label{eq:CE_expanded_2} \end{eqnarray} \end{widetext} Detailed expressions, with analytic formulae, for the average cluster functions of 3-body clusters as well as for the enthalpy of mixing represented as a function of average triple probabilities $y_n^{ijk}$, are given in Appendix A. \subsection{Chemical short-range order parameters} Short-range order in ternary alloys can be investigated by analyzing chemical pairwise interactions between unlike atoms. These pairwise interactions are related to $J_{|\omega|,n}^{(s)}$, where $|\omega|=2$ and $J_{|\omega|,n}^{(s)}$ are given by an inner product of the cluster function $\Gamma_{2,n}^{(s)}$ and the corresponding energy \cite{Asta1991,Wolverton1994}, namely \begin{equation} J_{2,n}^{(s)}=\langle \Gamma_{2,n}^{(s)}(\vec{\sigma}), E(\vec{\sigma}) \rangle= \rho_0^{(s)}\sum_{\left\{\vec{\sigma}\right\}}\Gamma_{2,n}^{(s)}(\vec{\sigma})E(\vec{\sigma}). \label{eq:ECI_def} \end{equation} Summation in the above equation is performed over all possible configurations and $\rho_0^{(s)}$ is a normalization constant chosen to satisfy the orthonormality criterion for cluster functions $\Gamma_{2,n}^{(s)}$. Effective cluster interactions in ternary alloys for pairs of non-zero point functions with indices (11),(12),(21) and (22) can now be written as \begin{equation} J_{2,n}^{(ij)}=\frac{4}{9}\sum_{h,k} E_n^{hk}\gamma_i(\sigma_h)\gamma_j(\sigma_k). \label{eq:ECI} \end{equation} where $E_n^{hk}$ is the average energy of configurations with atom $h$ being in the $n$-th nearest neighbour shell of atom $k$. From Eq. \ref{eq:ECI}, ECI for pairs with indices (11),(12),(21) and (22) are \begin{eqnarray} J_{2,n}^{(11)}&=&\frac{1}{9}\left(4E_n^{AA}+E_n^{BB}+E_n^{CC}-2E_n^{AB}-2E_n^{BA}\right. \nonumber \\ &-&\left.2E_n^{AC}-2E_n^{CA}+E_n^{BC}+E_n^{CB}\right), \nonumber \\ J_{2,n}^{(12)}&=&\frac{1}{2}\left(J_{2,n}^{(12)}+J_{2,n}^{(21)}\right)=\frac{\sqrt{3}}{9}\left(-E_n^{BB}+E_n^{CC}\right. \nonumber \\ &+&\left.E_n^{AB}+E_n^{BA}-E_n^{AC}-E_n^{CA}\right) \nonumber \\ J_{2,n}^{(22)}&=&\frac{1}{3}\left(E_n^{BB}+E_n^{CC}-E_n^{BC}-E_n^{CB}\right). \label{eq:ECI_pairs} \end{eqnarray} A chemical pairwise interaction between atoms $i$ and $j$ in the $n$-th neighbour shell in a ternary alloy is defined as the effective cluster interaction between pairwise clusters in binary alloys \cite{Asta1991,Wolverton1994} \begin{equation} V_n^{ij}=\frac{1}{4}\left(E_n^{ii}+E_n^{jj}-E_n^{ij}-E_n^{ji}\right), \label{eq:ECI_binary} \end{equation} where energies $E_n^{ii}$, $E_n^{jj}$, $E_n^{ij}$ and $E_n^{ji}$ are averaged over all the ternary alloy configurations. From Eqs. \ref{eq:ECI_pairs} and \ref{eq:ECI_binary}, a relation between chemical pairwise interactions involving unlike atoms, and effective cluster interactions of pairwise clusters in a ternary alloy, can be written in matrix form as \begin{equation} \left[\begin{array}{c} V_n^{AB} \\ V_n^{AC} \\ V_n^{BC} \\ \end{array} \right]= \left[ \begin{array}{ccc} \frac{9}{16} & \frac{-3\sqrt{3}}{8} & \frac{3}{16} \\ \frac{9}{16} & \frac{3\sqrt{3}}{8} & \frac{3}{16} \\ 0 & 0 & \frac{3}{4} \\ \end{array} \right] \left[\begin{array}{c} J_{2,n}^{(11)} \\ J_{2,n}^{(12)} \\ J_{2,n}^{(22)} \\ \end{array} \right]. \label{eq:VvsJ_Matrix} \end{equation} As for the binary alloy case, chemical pairwise interactions $V_n^{ij}$ have a simple meaning: $V_n^{ij}>0$ corresponds to attraction and $V_n^{ij}<0$ to repulsion between atoms $i$ and $j$. These interactions will be used in the analysis of SRO in Fe-Cr-Ni ternary alloys in Section IV.C. With Eq. \ref{eq:CE_expanded_2} expressed in terms of chemical pairwise interactions, the configurational enthalpy of mixing of a ternary alloy is given by \begin{widetext} \begin{eqnarray} \Delta H_{CE}(\vec{\sigma}) &=& J_1^{(0)}+J_1^{(1)}\left(1-3c_A\right) + J_1^{(2)}\frac{\sqrt{3}}{2}\left(c_C-c_B\right) \nonumber \\ &-&4\sum_{n}^{pairs} \left(V_n^{AB}y_n^{AB} + V_n^{AC}y_n^{AC}+V_n^{BC}y_n^{BC}\right) + \sum_{n}^{multibody} \ldots , \label{eq:CE_vs_V} \end{eqnarray} \end{widetext} SRO involving atoms $i$ and $j$ in the $n$-th nearest neighbour shell in either binary or ternary alloys can be described using the Warren-Cowley parameters $\alpha_n^{ij}$ \begin{equation} \alpha_n^{ij}=1-\frac{\left\langle p^{(i)},p^{(j)} \right\rangle_{n} }{\left\langle p^{(i)}\right\rangle \left\langle p^{(j)}\right\rangle} =1-\frac{y_n^{ij}}{c_ic_j}=1-\frac{P_n^{i-j}}{c_j}. \label{eq:SRO_definition} \end{equation} Here $n$ is a coordination sphere index, $c_i$ and $c_j$ are the concentrations of $i$'s and $j$'s atoms, and $P_n^{i-j}=y_n^{ij}/c_i$ is the conditional probability of finding atom $i$ in the $n$-th coordination sphere of atom $j$, see for example Ref. \onlinecite{DeFontaine1971}. As in the binary alloy case, $\alpha_n^{ij}$ vanishes if $P_n^{i-j}=c_j$, meaning that there is no (positive or negative) preference for a given atom to be surrounded by atoms of any other type. Segregation gives rise to positive $\alpha_n^{ij}$, whereas a negative value of $\alpha_n^{ij}$ indicates ordering. If at low concentration of atoms $j$, each atom $j$ is surrounded only by atoms $i$, i.e. ($P_n^{i-j}=1$), then $\alpha_n^{ij}$ acquires the lowest possible value $\alpha_{n,min}^{ij}=-(1-c_j)/c_j$. SRO parameters can be expressed in terms of average point and pair correlation functions. Inverting Eqs. \ref{eq:CE_M_Matrix}, \ref{eq:pair_cluster_funct} and \ref{eq:SRO_definition}, analytical formulae for SRO parameters in a ternary alloy become \begin{widetext} \begin{eqnarray} \alpha_n^{AB}&=&1-\frac{2-2\langle\gamma_1\rangle-2\sqrt{3}\langle\gamma_2\rangle-4\langle\gamma_1,\gamma_1\rangle_n+4\sqrt{3}\langle\gamma_1,\gamma_2\rangle_n}{2(1-2\langle\gamma_1\rangle)(1+\langle\gamma_1\rangle-\sqrt{3}\langle\gamma_2\rangle)} \nonumber \\ \alpha_n^{BC}&=&1-\frac{2+4\langle\gamma_1\rangle+2\langle\gamma_1,\gamma_1\rangle_n -6\langle\gamma_2,\gamma_2\rangle_n}{2(1+\langle\gamma_1\rangle-\sqrt{3}\langle\gamma_2\rangle)(1+\langle\gamma_1\rangle+\sqrt{3}\langle\gamma_2\rangle)} \nonumber \\ \alpha_n^{AC}&=&1-\frac{2-2\langle\gamma_1\rangle+2\sqrt{3}\langle\gamma_2\rangle-4\langle\gamma_1,\gamma_1\rangle_n - 4\sqrt{3}\langle\gamma_1,\gamma_2\rangle_n} {2(1-2\langle\gamma_1\rangle)(1+\langle\gamma_1\rangle+\sqrt{3}\langle\gamma_2\rangle)}. \label{eq:SRO_pairs} \end{eqnarray} \end{widetext} Since both point and pair correlation functions are generated by the ATAT package\cite{Walle2002} used in the present study, the SRO parameters of ternary alloys are going to be calculated using Eq. \ref{eq:SRO_pairs}. \subsection{Magnetic Cluster Expansion} Magnetic Cluster Expansion has been successfully applied to a number of binary systems, including bcc and fcc Fe-Cr \cite{Lavrentiev2010,Lavrentiev2011a} and fcc Fe-Ni \cite{Lavrentiev2014}. In MCE \cite{Lavrentiev2009,Lavrentiev2011}, each alloy configuration is defined by its chemical ($\sigma_i$) {\it and} magnetic ($\mathbf{M}_i$) degrees of freedom. MCE parameters are derived from DFT data on 30 ordered ternary Fe-Cr-Ni structures (see Supplementary Material), spanning the entire alloy composition range, together with DFT data on pure elements. Parametrization also used 29 binary fcc Fe-Ni configurations analysed in a recent application of MCE to fcc Fe-Ni alloys \cite{Lavrentiev2014}. We note that deriving exchange coupling parameters for non-collinear Hamiltonians from collinear \textit{ab initio} calculations is a known approach that provided a number of significant results for a broad variety of magnetic systems. This includes recent studies of MnSi by Hortamani {\it et al.} \cite{Hortamani2009}, and Fe$_{65}$Ni$_{35}$ by Liot and Abrikosov \cite{Liot2009}. Our own work on Fe and Fe/Cr interfaces \cite{Lavrentiev2010,Lavrentiev2011a}, which followed the same approach, agrees well with experiment and non-collinear {\it ab initio} calculations, thus further validating the above approach to the parametrization of Magnetic Cluster Expansion. To simplify applications of MCE to ternary alloys and reduce the number of fitting parameters, we use an MCE Hamiltonian that includes only pairwise interactions. In this approximation, the energy of an arbitrary alloy configuration $\left(\left\{\sigma_i\right\}, \left\{\mathbf{M}_i\right\} \right)$ is written in the Heisenberg-Landau form as \begin{widetext} \begin{eqnarray} H_{MCE}\left(\left\{\sigma_i\right\},\left\{\mathbf{M}_i\right\}\right)&=& \sum_i\mathcal{I}_{\sigma_i}^{(1)}+\sum_{ij\in 1NN}\mathcal{I}_{\sigma_i\sigma_j}^{(1NN)}+\sum_{ij\in 2NN}\mathcal{I}_{\sigma_i\sigma_j}^{(2NN)}+\ldots \nonumber \\ &+&\sum_iA_{\sigma_i}\mathbf{M}_i^2+\sum_iB_{\sigma_i}\mathbf{M}_i^4+\ldots \nonumber \\ &+&\sum_{ij\in 1NN}\mathcal{J}_{\sigma_i\sigma_j}^{(1NN)}\mathbf{M}_i\cdot \mathbf{M}_j+\sum_{ij\in 2NN}\mathcal{J}_{\sigma_i\sigma_j}^{(2NN)}\mathbf{M}_i\cdot \mathbf{M}_j+\ldots, \label{eq:MCE_Hamiltonian} \end{eqnarray} \end{widetext} where $\sigma_i, \sigma_j =$ Fe, Cr and Ni, and the non-magnetic and Heisenberg magnetic interaction parameters $\mathcal{I}_{ij}$ and $\mathcal{J}_{ij}$ for each coordination shell are represented by 3$\times$3 matrices. We take into account interactions that extend up to the fourth nearest neighbour coordination shell. Together, there are 24 independent non-magnetic and 24 independent magnetic interaction parameters. At the first stage of fitting, the on-site magnetic terms $A$, $B$, $C$,... were fitted using the energy versus magnetic moment curves computed for pure ferromagnetic Fe, Ni, and Cr. For chromium, only quadratic and quartic Landau expansion terms were used, while for iron and nickel the Landau expansion was extended to the 8$^{th}$-order in magnetic moment\cite{Lavrentiev2014}. The dependence of the on-site terms on atomic environment was neglected in order to reduce the number of parameters in the Hamiltonian. Following the methodology described in Ref. \cite{Lavrentiev2014}, the interaction terms $\mathcal{I}$ and $\mathcal{J}$ were fitted to DFT data on total energies and magnetic moments on each site in the simulation cell. Most of the alloy structures used for parameterizing the Fe-Cr-Ni MCE Hamiltonian (Eq. \ref{eq:MCE_Hamiltonian}) belong to the Fe-rich corner of the ternary alloy composition triangle. Hence we expect that MCE predictions are going to be most reliable for alloys where Fe content exceeds 50 at.\%. \subsection{Computational details} DFT calculations were performed using the Projector Augmented Wave (PAW) method implemented in VASP\cite{Kresse1996, Kresse1996a}. Exchange and correlation were treated in the generalized gradient approximation GGA-PBE \cite{Perdew1996}. To accelerate DFT calculations, we used PAW potentials without semi-core $p$ electron contribution. The core configurations of Fe, Cr and Ni in PAW potentials were [Ar]3d$^7$4s$^1$, [Ar]3d$^5$4s$^1$ and [Ar]3d$^9$4s$^1$, respectively. Total energies were calculated using the Monkhorst-Pack mesh\cite{Monkhorst1976} of $k$-points in the Brillouin zone, with $k$-mesh spacing of 0.2 $\AA^{-1}$. This corresponds to 14$\times$14$\times$14 or 12$\times$12$\times$12 $k$-point meshes for a two-atom bcc cubic cell or a four-atom fcc cubic cell, respectively. The plane wave cut-off energy used in the calculations was 400 eV. The total energy convergence criterion was set to 10$^{-6}$ eV/cell, and force components were relaxed to 10$^{-3}$ eV/$\AA$. Mapping DFT energies to CE was performed using the ATAT package\cite{Walle2002}. In order to find CE parameters for binary fcc alloys we used a database of 28 structures from Table I of Ref. \onlinecite{Barabash2006}. For binary bcc alloys we used the 58 structures from Table I of Ref. \onlinecite{Nguyen-Manh2007}. For ternary fcc alloys we used the 98 structures from Fig. 2 of Ref. \onlinecite{Garbulsky1994}. To our knowledge, there is no database of structures of ternary bcc alloys available at present. We constructed the input ternary bcc structures using binary structures of Ref. \onlinecite{Nguyen-Manh2007} as a starting point. The symmetry and the number of non-equivalent positions (NEPs) in each structure was checked, and structures for which the number of NEPs was greater than two were included in the ternary bcc structure database. The resulting input database for bcc ternary alloys consists of 94 structures. These structures are described in detail in Appendix B. Most of the collinear spin-polarized DFT calculations were performed assuming that the initial magnetic moments of Fe, Cr and Ni atoms were +3, -1 and +1 $\mu_B$, respectively. Since magnetic properties of Fe-Cr-Ni alloys are very complex in comparison with binary alloys, full relaxations starting from various initial magnetic configurations were performed in order to find the most stable magnetic order characterizing a given structure. Such an investigation was especially critical for fcc Fe-rich structures, where the energies of competing magnetic configurations are very close. Initial values of ECIs, derived by mapping to CE the DFT energies computed for the most stable magnetic configurations of input structures, provide a starting point for further refinement of CE parameters, which is performed by generating new structures. The complexity of magnetic properties of Fe-Cr-Ni alloys made it impossible to perform this refinement fully automatically, as is possible in the case of non-magnetic alloys. For example, the above choice of initial values of magnetic moments did not always lead to the most stable magnetic configurations. Hence results had to be filtered following an approach proposed in Ref. \onlinecite{Barabash2009}. For Fe-Cr-Ni alloys this meant that some of the structures had to be recalculated assuming an alternative initial magnetic configuration or, in a few extreme cases, the less stable structures were eliminated if their energies proved difficult to fit to a consistent set of ECIs. Despite the fact that performing fully automatic refinement of CE parameters was not possible, reasonable values of cross-validation error between DFT and CE formation enthalpies were achieved, proving that the final set of ECI describes interatomic interactions in Fe-Cr-Ni system fairly well. A detailed description of ECIs, the number of structures used in the fitting, and the cross-validation error between DFT and CE data is given in Section III. Quasi-canonical MC simulations were performed using the ATAT package \cite{Walle2002}. Most of the simulations were performed using a cell containing 8000 atoms in the form of 20$\times$20$\times$20 {\it primitive} fcc or bcc unit cells. For each composition, simulations were performed starting from a disordered high-temperature state (usually $T$ = 2500 K). The alloy was then cooled down with the temperature step of $\Delta T$ = 100K, with 5000 MC steps per atom at both thermalization and accumulation stages. Test simulations were also performed with 2000 MC steps at each of these stages. Since the results were not significantly different, there was no need to test with more than 5000 MC steps. A database of enthalpies of mixing and magnetic moments of ternary fcc Fe-Cr-Ni structures derived from DFT and used for fitting the MCE Hamiltonian (see Section V) is given in Supplementary Material. \section{Phase stability and magnetic properties at 0 K} \subsection{Pure Elements} \begin{table*} \caption{Volume per atom $V$, energy with respect to the energy of the ground state, $E-E_{GS}$, and magnetic moment per atom $|m_{tot}|$, computed for various structures of pure elements, compared to available experimental data. \label{tab:Fe}} \begin{ruledtabular} \begin{tabular}{cccccc} Struct. Name & $V$ (\AA$^3$/atom) & $V^{Expt.}$ (\AA$^3$/atom) & $E-E_{GS}$ (eV) & $|m_{tot}|$ ($\mu_B$) & $|m_{tot}^{Expt.}|$ ($\mu_B$) \\ \hline bcc-Fe (FM) -GS & 11.35 & 11.70\cite{Acet1994} & 0.000 & 2.199 & 2.22\cite{Crangle1963} \\ bcc-Fe (NM) & 10.46 & & 0.475 & 0.000 & \\ bcc-Fe (AFMSL) & 10.87 & & 0.444 & 1.290 & \\ bcc-Fe (AFMDL) & 11.34 & & 0.163 & 2.104 & \\ bcc-Fe (AFMTL) & 11.35 & & 0.112 & 4$\times$2.087; & \\ & & & & 2$\times$2.351 & \\ fcc-Fe (NM) & 10.22 & & 0.167 & 0.000 & \\ fcc-Fe (FM-HS) & 11.97 & 12.12\cite{Acet1994} & 0.153 & 2.572 & \\ fcc-Fe (FM-LS) & 10.52 & & 0.162 & 1.033 & \\ fcc-Fe (AFMSL) & 10.76 & 11.37\cite{Acet1994} & 0.100 & 1.574 & 0.75\cite{Abrahams1962} \\ fcc-Fe (AFMDL) & 11.20 & & 0.082 & 2.062 & \\ fcc-Fe (AFMTL) & 11.45 & & 0.082 & 8$\times$2.155; & \\ & & & & 4$\times$2.429 & \\ \\ \hline bcc-Cr (AFMSL) -GS & 11.63 & 11.94\cite{Kittel1971}& 0.000 & 1.070 & \\ bcc-Cr (NM) & 11.41 & & 0.011 & 0.000 & \\ fcc-Cr (NM) & 11.75 & & 0.405 & 0.000 & \\ \\ \hline fcc-Ni (FM) -GS & 10.91 & 10.90 \cite{Kittel1971} & 0.000 & 0.641 & 0.60\cite{Crangle1963}\\ fcc-Ni (NM) & 10.84 & & 0.056 & 0.000 & \\ bcc-Ni (FM) & 11.00 & & 0.092 & 0.569 & \\ bcc-Ni (NM) & 10.90 & & 0.107 & 0.000 & \\ \end{tabular \end{ruledtabular} \end{table*} Magnetism of Fe-Cr-Ni alloys gives rise to several structural and magnetic instabilities. This effect is well known in pure iron. \textit{Ab initio} analysis of structural and magnetic phase stability of iron was performed in Refs. \onlinecite{Herper1999,Moruzzi1989}. Our calculations confirm that the most stable Fe phase at 0 K is the ferromagnetic (FM) bcc phase. Anti-ferromagnetic single layer (AFMSL) and anti-ferromagnetic double layer (AFMDL) fcc structures are more stable than the high-spin (HS) and low-spin (LS) ferromagnetic configurations. We have extended analysis of anti-ferromagnetism in iron to anti-ferromagnetic triple layer (AFMTL) fcc and bcc structures. We have found that fcc-Fe AFMTL has the same energy per atom as fcc-Fe AFMDL but they have significantly different volumes, see Table \ref{tab:Fe}. The bcc Fe AFMTL structure of iron is more stable than bcc Fe AFMSL and bcc Fe AFMDL, but it is still less stable than bcc Fe FM. DFT calculations confirm that the most stable collinear magnetic Cr and Ni phases at 0 K are anti-ferromagnetic bcc and ferromagnetic fcc. Ferromagnetic bcc Ni and non-magnetic fcc Cr are 0.096 eV/atom and 0.405 eV/atom less stable than fcc Ni and bcc Cr, respectively. Since the ground states of Fe, Cr and Ni belong to different crystal lattices, the phase stability of Fe-Cr-Ni alloys and binary sub-systems is analyzed in terms of their enthalpies of formation, defined as the energy of the alloy, calculated at zero pressure, with respect to the energies of ferromagnetic bcc-Fe, ferromagnetic fcc-Ni, and anti-ferromagnetic bcc-Cr. To investigate properties of alloys on fcc and bcc crystal lattices, stabilities of fcc and bcc alloys have also been analyzed in terms of their enthalpy of mixing, defined as the energy of an alloy with respect to the energies of fcc or bcc structures of pure elements, where the choice of bcc or fcc depends on the choice of the crystal structure of the alloy under consideration. \subsection{Fe-Ni binary alloys} \begin{figure*} \includegraphics[width=\linewidth]{Fe-Ni_binary_results} \caption{ (Color online) Enthalpies of mixing (a,b), volumes per atom (c,d) and magnetic moments (e,f) of Fe-Ni structures on fcc (a,c,e) and bcc (b,d,f) lattices, calculated using DFT. Experimental data are taken from Refs. \onlinecite{Landolt,Crangle1963,Chamberod1979}. GS refers to the ground state on fcc (a,c) or bcc (b,d) crystal lattices; ST is the most stable structure and magnetic configuration for the corresponding alloy composition.} \label{fig:FeNi_results} \end{figure*} There is extensive literature on models for Fe-Ni alloys, see for example Refs. \onlinecite{Barabash2009,Mohri2004,Tucker2008,Abrikosov2007,Crisan2002,Ekholm2010,Ruban2005,Ruban2007}. Recently \cite{Lavrentiev2014} we used a DFT database to parameterize the Magnetic Cluster Expansion and to investigate magnetic properties of Fe-Ni alloys. In this sub-section, we compare our DFT results with previous experimental and theoretical studies, focusing on the stability of magnetic configurations and on equilibrium volumes of alloy structures. Our results agree with an assertion, derived from simulations \cite{Barabash2009,Mohri2004,Tucker2008} and experiments \cite{Massalski1990,Reuter1989}, that fcc FeNi (L1$_0$), FeNi$_3$ (L1$_2$) and FeNi$_8$ (Pt$_8$Ti-like\cite{Barabash2009}) compounds are the global (on both fcc {\it and} bcc lattices) alloy ground states for the relevant compositions, see Fig. \ref{fig:FeNi_results}(a). Our results agree with Ref. \onlinecite{Barabash2009} in that the fcc ferromagnetic Z1(100) phase of Fe$_3$Ni (see Fig. 3 in Ref. \onlinecite{Barabash2009}) is more stable than L1$_2$, contrary to what was previously assumed according to Refs. \onlinecite{Massalski1990,Crisan2002,Mohri2004,Mohri2009}. In Ref. \onlinecite{Barabash2009} the AFMDL configuration of fcc Fe-Ni alloys was not investigated, despite the fact that AFMDL represents the most stable magnetic configuration of fcc-Fe, see our Table \ref{tab:Fe} and Refs. \onlinecite{Herper1999,Moruzzi1989,Klaver2012}. In relation to the AFMDL structure of fcc Fe-Ni, the Z1 Fe$_3$Ni structure\cite{Lu1991} does not represent the ground state, and instead an alternative fcc ground state, Fe$_3$Ni$_2$ with $I4/mmm$ symmetry, is predicted by CE, see Fig. \ref{fig:FeNi_results}(a). None of the AFM fcc structures is the actual ground state, however the energies of fcc Fe$_5$Ni AFMTL, ferri-magnetic fcc Fe$_5$Ni, and fcc Fe$_4$Ni AFMTL, are fairly close to the bottom of the zero temperature phase stability curve. The existence of these magnetic structures may affect finite temperature stability of fcc alloys. Our CE calculations also predict two bcc ground states, Fe$_4$Ni$_5$ (VZn-like\cite{Nguyen-Manh2007}) and FeNi$_5$ (of $Cmmm$ symmetry) that are still less stable than fcc structures of similar compositions, see Figs. \ref{fig:FeNi_results}(b) and \ref{fig:formation_binaries}. Fe$_4$Ni$_5$ (VZn-like) bcc structure is predicted as the lowest energy alloy configuration by both DFT and CE simulations. Enthalpies of mixing of fcc and bcc Fe-Ni structures calculated using DFT and CE are compared in Fig. \ref{fig:FeNi_results}(a,b). To remain consistent with the treatment of binary alloys Fe-Cr and Ni-Cr, we used the same sets of cluster interaction parameters, namely five two-body, three three-body, two four-body, one five-body clusters, for fcc binary alloys, and five two-body, two three-body, one four-body, one five-body clusters for the corresponding bcc alloys. A set of ECIs obtained by mapping energies of structures from DFT to CE is given in Fig. \ref{fig:ECI_binaries} and Table \ref{tab:ECI_binary} in Appendix C. The cross-validation errors between DFT and CE are 8.1 and 10.9 meV/atom for fcc and bcc Fe-Ni alloys, respectively. The magnitude and sign of ECIs explain the behaviour of fcc and bcc Fe-Ni alloys found in simulations. In fcc alloys the first and third nearest neighbour (1NN and 3NN) pair interactions are positive, whereas the second nearest neighbour (2NN) interaction is negative. In binary alloys, from Eqs. \ref{eq:CE_1} and \ref{eq:CE_2}, this favours having the unlike atoms occupying the first and the third neighbour coordination shell, and the like atoms occupying the second neighbour shell. For the fcc lattice this favours the formation of L1$_2$ intermetallic phase, which is the ground state of fcc Fe-Ni alloy. In bcc alloys the 1NN Fe-Ni pair interaction is negative, corresponding to repulsive interaction between the unlike atoms in the first neighbour shell. The 2NN pair interaction is positive and similar in its magnitude to the 1st ECI. As a result, bcc Fe-Ni alloys exhibit several intermetallic phases with negative enthalpies of mixing. The atomic volumes of fcc and bcc alloys shown in Fig. \ref{fig:FeNi_results}(c,d) are not linear functions of Ni content. This non-linearity stems from the difference between atomic sizes of Fe and Ni {\it and} magnetism, see Fig. \ref{fig:FeNi_results}(e,f). Bcc alloys with low Ni content have larger volume per atom than pure Fe, despite the fact that Ni atoms have smaller size. This is correlated with the fact that the Fe$_{15}$Ni structure has the largest average atomic magnetic moment, 2.31 $\mu_B$. In fcc Fe-Ni alloy the non-linearity of atomic volume as a function of Ni content is even more pronounced, since alloys with Ni content lower than 25\% exhibit anti-ferromagnetic interaction between Fe and Ni, resulting in higher atomic density than ferromagnetically ordered alloys. Experimental measurements \cite{Chamberod1979} show that the average atomic volume is maximum for Fe-Ni alloys with $\sim 37$ at. \% Ni. This is correlated with the fact that the Fe$_3$Ni$_2$ intermetallic phase has the largest volume per atom, see Fig.\ref{fig:FeNi_results}(c). There are several structures with smaller Ni content that are ferromagnetically ordered at 0K and have larger volumes per atom than Fe$_3$Ni$_2$. Those structures are metastable, and alloys with Ni concentration below 40 at. \% Ni are mixtures of ferromagnetic Fe$_3$Ni$_2$, anti-ferromagnetic Fe, and metastable ferromagnetic and anti-ferromagnetic alloy phases. Near 25 at. \% Ni concentration the most stable magnetic configurations are ferromagnetic, however the energy difference between them and anti-ferromagnetic phases, characterized by smaller volumes, is fairly small. In particular, the most stable structure corresponding to 33 at. \% Ni is a ferromagnetic $\beta$-phase\cite{Barabash2009} where the enthalpy of mixing is -0.070 eV/atom and the atomic volume is 11.47 $\AA^3$ per atom. The AFMTL structure is 0.023 eV/atom less stable, and has the atomic volume of 11.39 $\AA^3$, whereas AFMSL is 0.039 eV/atom less stable than FM and has the volume of 11.24 $\AA^3$ per atom. The coexistence of structures with different magnetic order and different atomic volumes but similar energies is the origin of the Invar effect \cite{Entel1993}. \subsection{Fe-Cr binary alloys} \begin{figure*} \includegraphics[width=\linewidth]{Fe-Cr_binary_results} \caption{ (Color online) Enthalpies of mixing (a,b), volumes per atom (c,d) and magnetic moments (e,f) predicted by DFT for fcc (a,c,e) and bcc (b,d,f) Fe-Cr alloys. Experimental data are taken from Refs. \onlinecite{Aldred1976,Aldred1976a,Kajzar1980}. GS - ground states of alloys on fcc (a,c) or bcc (b,d) crystal lattices, ST - the most stable structure for a given composition.} \label{fig:FeCr_results} \end{figure*} Extensive theoretical \cite{Nguyen-Manh2007,Nguyen-Manh2008,Nguyen-Manh2009,Nguyen-Manh2012,Lavrentiev2007,Lavrentiev2010,Olsson2003,Olsson2006,Zhang2009,Erhart2008} and experimental\cite{Mirebeau1984} investigations show that low Cr bcc Fe-Cr alloys form intermetallic phases where the most stable structures contain between 6.25 and 7.41 at. \% Cr \cite{Nguyen-Manh2007,Erhart2008}. Results of calculations shown in Fig. \ref{fig:FeCr_results}(b) confirm those findings. For fcc Fe-Cr alloys, we predict three new ground states: Fe$_3$Cr(L1$_2$), FeCr$_2$($\beta2$(100)\cite{Barabash2009}) and FeCr$_8$ (Pt$_8$Ti-like) that are all significantly less stable than bcc structures, see Figs. \ref{fig:FeCr_results}(a) and \ref{fig:formation_binaries}. Enthalpies of mixing of ordered Fe$_3$Cr and FeCr$_2$ structures are -0.111 and -0.120 eV/atom, and are approximately 0.05 eV/atom lower than those calculated for fcc Fe-Cr random alloys. Comparison between enthalpies of mixing of fcc and bcc Fe-Cr alloys calculated using DFT and CE is shown in Figs. \ref{fig:FeCr_results}(a) and \ref{fig:FeCr_results}(b). A full set of ECIs derived by mapping DFT energies to CE is given in Fig. \ref{fig:ECI_binaries}(c,d) and Table \ref{tab:ECI_binary} in Appendix C. The cross-validation error between DFT and CE is 11.3 and 10.6 meV/atom for fcc and bcc Fe-Cr alloys, respectively. Similarly to fcc Fe-Ni alloys, the first and the third nearest neighbour (1NN and 3NN) pair interactions are positive and the second nearest neighbour (2NN) interaction is negative, favouring the L1$_2$ intermetallic phase, which is also the ground state of fcc Fe-Cr alloy. The 1NN pair interaction in bcc Fe-Cr alloys is negative, as in bcc Fe-Ni alloys, implying repulsive interaction between the unlike atoms in the first nearest neighbour coordination shell. ECIs of bcc Fe-Cr alloys were previously analyzed in Ref. \onlinecite{Lavrentiev2007}. Despite the fact that our DFT calculations use a different set of clusters, our results are in agreement with Ref. \onlinecite{Lavrentiev2007} in that the dominant negative 1NN pair interaction and positive fifth nearest neighbour pair interaction together give rise to the formation of Fe - 6.25 at.\% Cr $\alpha$-phase. Atomic volumes of bcc Fe-Cr alloys remain nearly constant over a broad range of alloy compositions, exhibiting small variation in the interval of 0.3 $\AA{}^3$ per atom, see Fig.\ref{fig:FeCr_results}(d). There are two exceptions to this rule. The volume per atom in Cr-rich alloys decreases as a function of Fe content. This can be explained by the fact that Fe impurities interfere with anti-ferromagnetic ordering of magnetic moments in pure Cr, reducing the magnitude of moments and the strength of magnetic interactions, see Fig. \ref{fig:FeCr_results}(f). This also affects the average atomic volume. In Fe-rich alloys, atomic volume increases linearly with Cr content, reaching a maximum of 11.50 $\AA^3$ per atom at 8.33 at. \% Cr. This confirms previous theoretical predictions derived using CPA and SQS methods \cite{Olsson2006,Zhang2009}, which show a local maximum of atomic volume (lattice parameter) in random bcc Fe-Cr alloys at approximately 10 at. \% Cr. These theoretical predictions are in agreement with experimental data\cite{Pearson1958}, where the observed deviation from Vegard's law is largest at $\sim$ 10 at. \% Cr . This effect probably results from magneto-volume coupling and strong anti-ferromagnetic interaction between Fe and Cr atoms. At low density magnetic moments are larger and the energy of atomic structure is lower, hence Cr impurities in Fe tend to increase volume per atom in the $\alpha$-phase. The increase is almost linear in Cr content until a critical concentration is reached and Cr starts segregating. At variance with DFT analysis of ordered structures performed here, and earlier studies of random alloys \cite{Olsson2006,Zhang2009}, the experimentally measured atomic volume in alloys with Cr concentration higher than 10 \% continues to increase linearly towards the limit of pure Cr. The likely reason for the lack of agreement between DFT and experiment is that neither the ordered structures treated here nor the random alloys investigated in Refs. \onlinecite{Olsson2006,Zhang2009} are representative of real bcc Fe-Cr alloys, where alloy microstructure is a mixture of $\alpha$-phase and Cr clusters, as shown in Figs. \ref{fig:FeCr_results}(b,d) by black circles\cite{Lavrentiev2007}. The composition dependence of atomic volume in fcc Fe-Cr alloys differs significantly from what is found in fcc Fe-Ni alloys. Due to strong anti-ferromagnetic interaction between Fe and Cr atoms, anti-ferromagnetic or ferri-magnetic order dominates in the entire range of alloy compositions, see Fig. \ref{fig:FeCr_results}(e). Volume decrease caused by anti-ferromagnetic ordering in Fe-rich fcc Fe-Ni alloys is also present in the entire range of alloy compositions. Volume decrease as a function of Cr concentration is particularly strongly pronounced in Cr-rich fcc Fe-Cr alloys. Magnetic moments of Fe and Cr atoms as well as the average magnetic moment of ordered bcc Fe-Cr structures are similar to those predicted for random alloys in Refs. \onlinecite{Olsson2006,Klaver2006}. They agree well with the available experimental data \cite{Aldred1976,Aldred1976a,Kajzar1980}. \subsection{Cr-Ni binary system} \begin{figure*} \includegraphics[width=\linewidth]{Cr-Ni_binary_results} \caption{ (Color online) Enthalpies of mixing (a,b), volumes per atom (c,d) and magnetic moments (e,f) calculated using DFT for Cr-Ni alloys on fcc (a,c,e) and bcc (b,d,f) lattices. Experimental data are taken from Ref. \onlinecite{Landolt}. GS - ground states of alloys on fcc (a,c) or bcc (b,d) lattices, ST - the most stable structure found for a given alloy composition.} \label{fig:CrNi_results} \end{figure*} DFT and CE simulations of fcc Cr-Ni alloys were performed in Ref. \onlinecite{Tucker2008}. Our analysis confirms the conclusion, derived from simulations and experiment, that there is only one globally stable ground state of the alloy, realized on the CrNi$_2$ (MoPt$_2$-like) ordered structure. We find a further five fcc ground states: Cr$_7$Ni (of $Cmmm$ symmetry, predicted by CE), Cr$_5$Ni (also predicted by CE, with $Cmmm$ symmetry), Cr$_3$Ni-Z1(100), Cr$_5$Ni$_2$ (of $I4/mmm$ symmetry, also predicted by CE), and Cr$_2$Ni-$\beta1$(100). The last of these is characterized by a large positive value of the enthalpy of formation, and is less stable than bcc alloys with the same composition, see Figs. \ref{fig:CrNi_results}(a) and \ref{fig:formation_binaries}. We find only one alloy configuration on a bcc lattice that has small negative enthalpy of mixing, CrNi (predicted by CE, with $Cmmn$ symmetry and $H_{mix}=-4$ meV/atom). Comparison of enthalpies of mixing of fcc and bcc Cr-Ni alloys calculated using DFT and CE is given in Fig. \ref{fig:CrNi_results}(a,b). A full set of ECIs found by mapping the energies of structures from DFT to CE is given in Fig. \ref{fig:ECI_binaries} and Table \ref{tab:ECI_binary} in Appendix C. Cross-validation errors between DFT and CE are 14.2 and 12.8 meV/atom for fcc and bcc Cr-Ni alloys, respectively. Similarly to fcc Fe-Ni and Fe-Cr alloys, the first and third nearest neighbour (1NN and 3NN) pair interactions in fcc Cr-Ni alloys are positive and the second nearest neighbour (2NN) interaction is negative. Unlike the other two binary systems, the ground state of fcc Cr-Ni alloys is MoPt$_2$-like phase. The ECI parameters derived from our DFT calculations and the cross-validation error between DFT and CE are in agreement with those of Ref. \onlinecite{Tucker2008}. The negative 1NN pair interaction in bcc Cr-Ni system is the largest of all the binary alloys. Because of that, there is only one bcc intermetallic phase, CrNi, of \textit{Cmmn} symmetry, which has small negative enthalpy of mixing ($-4$ meV/atom). Variation of atomic volume as a function of Ni content in both fcc and bcc alloys is more linear than in Fe-Ni and Fe-Cr alloys because magnetic interactions are weaker, see Figs. \ref{fig:CrNi_results}(c-f). Similarly to Fe-Cr alloys, the difference between atomic volumes of alloys with low and high concentration of Cr is more significant in fcc than bcc alloys. \begin{figure*} \centering \begin{minipage}{.50\textwidth} \centering a) \includegraphics[width=.9\linewidth]{ECI_FeNi_fcc} \end{minipage \begin{minipage}{.50\textwidth} \centering b) \includegraphics[width=.9\linewidth]{ECI_FeNi_bcc} \end{minipage} \newline \begin{minipage}{.50\textwidth} \centering c) \includegraphics[width=.9\linewidth]{ECI_FeCr_fcc} \end{minipage \begin{minipage}{.50\textwidth} \centering d) \includegraphics[width=.9\linewidth]{ECI_FeCr_bcc} \end{minipage} \newline \begin{minipage}{.50\textwidth} \centering e) \includegraphics[width=.9\linewidth]{ECI_CrNi_fcc} \end{minipage \begin{minipage}{.50\textwidth} \centering f) \includegraphics[width=.9\linewidth]{ECI_CrNi_bcc} \end{minipage} \caption{ Effective cluster interactions (ECIs) derived using CE method for fcc Fe-Ni (a), bcc Fe-Ni (b), fcc Fe-Cr (c), bcc Fe-Cr (d),fcc Cr-Ni (e), bcc Cr-Ni (f) alloys. } \label{fig:ECI_binaries} \end{figure*} \subsection{Fe-Cr-Ni ternary system} The stability of fcc and bcc phases of ternary Fe-Cr-Ni alloys, and the corresponding binary alloys, is defined with respect to bcc Fe, bcc Cr and fcc Ni, as mentioned previously. Enthalpies of formation of Fe-Ni, Fe-Cr and Cr-Ni alloys are shown in Figs. \ref{fig:formation_binaries}(a), \ref{fig:formation_binaries}(b) and \ref{fig:formation_binaries}(c), respectively. The Ni-rich fcc Fe-Ni and Cr-Ni alloys are usually more stable than bcc alloys of similar composition, whereas alloys with smaller Ni content tend to adopt bcc structure. In Fe-Cr alloys, energies of fcc phases are always higher than the energies of bcc structures. Even so, meta-stable fcc Fe-Cr structures and interactions between the unlike atoms in fcc Fe-Cr alloys prove critical to understanding chemical ordering in Fe-Cr-Ni alloy system. From the list of ground states associated with each lattice type shown in Figs. \ref{fig:FeNi_results}, \ref{fig:FeCr_results} and \ref{fig:CrNi_results}, we conclude that there are only four binary fcc phases: FeNi, FeNi$_3$, FeNi$_8$ and CrNi$_2$, and only one binary bcc Fe-Cr phase, namely the $\alpha$-phase, which are the global ground states of the alloys. Enthalpies of formation, volumes and magnetic moments per atom, and space groups of the relevant alloy structures are given in Table \ref{tab:Enthalpies_of_GS}. \begin{figure*} \centering \begin{minipage}{.50\textwidth} \centering a)\includegraphics[width=.9\linewidth]{formation_FeNi} \end{minipage \begin{minipage}{.50\textwidth} \centering b)\includegraphics[width=.9\linewidth]{formation_FeCr} \end{minipage} \newline \begin{minipage}{.50\textwidth} \centering c)\includegraphics[width=.9\linewidth]{formation_CrNi} \end{minipage} \caption{ Enthalpies of formation of Fe-Ni (a), Fe-Cr (b), and Cr-Ni (c) binary structures.} \label{fig:formation_binaries} \end{figure*} Enthalpies of formation of fcc and bcc Fe-Cr-Ni alloys derived from DFT and CE are compared in Fig. \ref{fig:phase_stability_FeCrNi}. The most stable fcc and bcc structures form convex hulls, shown in Fig. \ref{fig:phase_stability_FeCrNi} by blue and red surfaces, respectively. The line of intersection between these two surfaces corresponds to the zero Kelvin fcc-bcc phase transition, which occurs if the enthalpies of formation of bcc an fcc alloys are equal. There is no Fe-Cr-Ni ternary alloy configuration on bcc lattice that has negative enthalpy of formation. Fcc alloy structures have negative enthalpy of formation in the Ni-rich limit of alloy compositions. This region of negative enthalpy of formation is elongated along the Fe-Ni edge of the alloy composition triangle. The L1$_2$-based fcc Fe$_2$CrNi phase, similar to Cu$_2$NiZn alloy phase, is the global ground state of Fe-Cr-Ni alloys. The enthalpy of formation, volume per atom, magnetic moments of each atom, as well as the space group of Fe$_2$CrNi structure, are given in Table \ref{tab:Enthalpies_of_GS}. ECIs of ternary fcc and bcc alloys are derived by mapping DFT energies onto CE for 248 fcc and 246 bcc structures, respectively. In CE simulations we used the same set of clusters as in fcc (five two-body, three three-body, two four-body, one five-body clusters) and bcc (five two-body, two three-body, one four-body, one five-body clusters) binary alloys. Since in ternary alloys each cluster can be decorated by point functions in various ways (see Section II.A and Table I), the number of ECIs is much larger than the number of clusters taken into consideration. Namely, we have 15 two-body, 16 three-body, 14 four-body, 12 five-body clusters for fcc alloys and 15 two-body, 12 three-body, 6 four-body, 18 five-body clusters for bcc alloys. Values of all the optimized ECIs for ternary alloys are given in Fig. \ref{fig:ECI_FeCrNi_ternary} and Table \ref{tab:ECI_def_ternary}. Cross-validation errors between DFT and CE are 10.2 and 11.2 meV/atom for fcc and bcc ternary alloys, respectively. \begin{table*} \caption{Enthalpies of formation of the lowest energy intermetallic phases of fcc Fe-Cr-Ni ternary alloys. \label{tab:Enthalpies_of_GS}} \begin{ruledtabular} \begin{tabular}{cccccccc} Structure & Space & Wyckoff & Mag. space & (Mag.) Wyckoff & $V$ & $H_{form}$ & $M$ \\ & group& positions & group & positions & (eV) & $H_{form}$ & ($\mu_B$) \\ \hline FeNi & $P4/mmm$ & Fe$_1$ 2$e$ & $P4/mm'm'$ & Fe$_1$ 2$e$ & 11.33 & -0.069 & 2.66 \\ (L1$_0$) & & Ni$_1$ 1$a$ & & Ni$_1$ 1$a$ & & & 0.63 \\ & & Ni$_2$ 1$c$ & & Ni$_2$ 1$c$ & & & 0.63 \\ FeNi$_3$ & $Pm-3m$ & Fe$_1$ 1$a$ & $Pm'm'm$ & Fe$_1$ 1$a$ & 11.13 & -0.091 & 2.91 \\ (L1$_2$) & & Ni$_1$ 3$c$ & & Ni$_1$ 1$f$ & & & 0.59 \\ & & & & Ni$_2$ 1$d$ & & & 0.58 \\ & & & & Ni$_3$ 1$g$ & & & 0.72 \\ FeNi$_8$ & $I4/mmm$ & Fe$_1$ 2$a$ & $P-1$ & Fe$_1$ 1$a$ & 10.98 & -0.051 & 2.81 \\ (NbNi$_8$) & & Ni$_1$ 8$h$ & & Ni$_1 $2$i$ & & & 0.60 \\ & & Ni$_2$ 8$i$ & & Ni$_2$ 2$i$ & & & 0.63 \\ & & & & Ni$_3$ 2$i$ & & & 0.61 \\ & & & & Ni$_4$ 2$i$ & & & 0.61 \\ CrNi$_2$ & $Immm$ & Cr$_1$ 2$a$ & $Immm1'$ & Cr$_1$ 2$a$ & 10.91 & -0.016 & 0.00 \\ (MoPt$_2$) & & Ni$_1$ 4$e$ & & Ni$_1$ 4$e$ & & & 0.00 \\ Fe$_2$CrNi & $P4/mmm$ & Cr$_1$ 1$c$ & $Pm'm'm$ & Cr$_1$ 1$f$ & 11.37 & -0.026 & -2.44 \\ (Cu$_2$NiZn) & & Fe$_1$ 2$e$ & & Fe$_1$ 1$d$ & & & 2.05 \\ & & Ni$_1$ 1$a$ & & Fe$_2$ 1$g$ & & & 2.12 \\ & & & & Ni$_1$ 1$a$ & & & 0.15 \\ \end{tabular \end{ruledtabular} \end{table*} \begin{figure} \includegraphics[width=\columnwidth]{BCC_FCC_stability} \caption{(Color online) Enthalpies of formation predicted by DFT (filled circles) and CE (open circles) for ternary Fe-Cr-Ni alloys at 0K. Only the most stable structures for each composition are shown. Blue and red circles show computed formation enthalpies of fcc and bcc Fe-Cr-Ni ternary alloys. Blue and red surfaces show convex hulls for fcc and bcc crystal structures, respectively. Black solid line corresponds to the intersection between fcc and bcc convex hulls. Cross-validation errors between DFT and CE are 10.2 and 11.2 meV/atom for fcc and bcc ternary alloys, respectively. \label{fig:phase_stability_FeCrNi}} \end{figure} \begin{figure*} \centering \begin{minipage}{\textwidth} \centering a)\includegraphics[width=0.9\textwidth]{ECI_FeCrNi_ptqp_fcc} \end{minipage \newline \begin{minipage}{\textwidth} \centering b)\includegraphics[width=0.9\textwidth]{ECI_FeCrNi_ptqp_bcc} \end{minipage \caption{ Effective cluster interactions obtained using the CE method for fcc (a) and bcc (b) Fe-Cr-Ni ternary alloys. \label{fig:ECI_FeCrNi_ternary}} \end{figure*} Volumes per atom of fcc and bcc Fe-Cr-Ni ternary alloy structures computed using DFT at 0K are shown in Fig. \ref{fig:volumes_FeCrNi}. Both fcc and bcc alloy configurations exhibit the largest volume per atom in the Cr-rich corner of the diagram. Atomic volume is smallest in the Ni-rich corner. The difference between the two values is larger for fcc alloys. Atomic volumes of fcc structures exhibit a significant degree of non-linearity as functions of alloy composition. This is explained by different magnetic behaviour of fcc and bcc alloys, see Figure \ref{fig:mag_FeCrNi}(a-d), treated as a function of alloy composition. A relation between fcc-bcc phase stability and magnetic moments of the most stable structures, as well as the discontinuity in the magnitude of the average magnetic moment at the fcc-bcc phase transition line, are illustrated in Fig. \ref{fig:mag_FeCrNi}(e). Average magnetic moments in bcc alloys are almost linear functions of Fe content. Magnetic moments are maximum for the Fe-rich alloy compositions and minimum for the anti-ferromagnetically ordered Cr-rich alloys. Fcc Fe-rich alloys do not exhibit large average magnetic moments, which are ordered anti-ferromagnetically, similarly to Fe-Ni alloys discussed in Section III.B. Alloys corresponding to the centre of the composition triangle, characterized by the approximately equal amounts of Fe, Cr and Ni, have relatively small average magnetic moments. The average magnetic moment decreases rapidly with increasing Cr content. For example, the average atomic magnetic moment in fcc (Fe$_{0.5}$Ni$_{0.5}$)$_{1-x}$Cr$_x$ alloys is 1.63, 0.97, 0.69 and 0.00 $\mu_B$ for Cr content $x$ = 0.0, 0.2, 0.33 and 0.5, respectively. These results are in agreement with experimental observations, performed at 4.2K, and showing that magnetization decreases rapidly in Fe$_{0.65}$(Cr$_x$Ni$_{1-x}$)$_{0.35}$ alloys as a function of Cr content in the interval from $x$ = 0.0 to 0.2 \cite{Rode1976}. This effect is also responsible for the observed reduction of the Curie temperature as a function of Cr content in Fe$_{0.65}$(Cr$_x$Ni$_{1-x}$)$_{0.35}$ and (Fe$_{0.5}$Ni$_{0.5}$)$_{1-x}$Cr$_x$ alloys, described in Refs. \onlinecite{Rode1976} and \onlinecite{Bansal1976}. Non-linear variation of magnetic moments as functions of alloy composition in fcc alloys results in deviations from Vegard's law. Despite the fact that Cr atoms have larger size, the volume per atom of fcc (FeNi)$_{1-x}$Cr$_x$ alloys decreases as a function of $x$, and is 11.33, 11.20, 11.09 and 10.92 $\AA^3$/atom for $x$ = 0.0, 0.2, 0.33 and 0.5, respectively. Results for other compositions are given in Supplementary Material. \begin{figure*} \centering \begin{minipage}{.50\textwidth} \centering a)\includegraphics[width=.9\linewidth]{volume_fcc_paper_hot} \end{minipage \begin{minipage}{.50\textwidth} \centering b)\includegraphics[width=.9\linewidth]{volume_bcc_paper_hot} \end{minipage} \newline \begin{minipage}{.50\textwidth} \centering c)\includegraphics[width=.9\linewidth]{volume_fcc_2D_hot} \end{minipage \begin{minipage}{.50\textwidth} \centering d)\includegraphics[width=.9\linewidth]{volume_bcc_2D_hot} \end{minipage} \caption{(Color online) Volumes (in $\AA^3$) of stable fcc (a, c) and bcc (b, d) ordered structures predicted by DFT calculations at 0K for various alloy compositions. Filled and open circles in (a, b) correspond to DFT data above and below the interpolated values, represented by the respective surfaces. (c, d) are the orthogonal projections of (a, b). } \label{fig:volumes_FeCrNi} \end{figure*} \begin{figure*} \centering \begin{minipage}{.50\textwidth} \centering a)\includegraphics[width=.9\linewidth]{mag_fcc_paper_hot} \end{minipage \begin{minipage}{.50\textwidth} \centering b)\includegraphics[width=.9\linewidth]{mag_bcc_paper_hot} \end{minipage} \newline \begin{minipage}{.50\textwidth} \centering c)\includegraphics[width=.9\linewidth]{mag_fcc_2D_hot} \end{minipage \begin{minipage}{.50\textwidth} \centering d)\includegraphics[width=.9\linewidth]{mag_bcc_2D_hot} \end{minipage} \newline \begin{minipage}{.50\textwidth} \centering e)\includegraphics[width=.9\linewidth]{mag_fccVSbcc_2D_hot} \end{minipage} \caption{(Color online) Magnetic moment per atom (in $\mu_B$) in the most stable fcc (a, c) and bcc (b, d) ordered alloy structures predicted by DFT calculations at 0K for each alloy composition. Filled and open circles in (a, b) correspond to DFT data above and below the interpolated values represented by the respective surfaces. (c, d) are the orthogonal projections of (a, b). (e) The average atomic magnetic moment of fcc and bcc structures is discontinuous across the fcc-bcc phase transition line, shown in the figure as solid black line. } \label{fig:mag_FeCrNi} \end{figure*} Magnetic moments of each component of fcc and bcc alloys are shown in Fig. \ref{fig:mag_FeCrNi_components}. The results exhibit a rapid decrease of magnetic moments on Ni sites as functions of Cr content in fcc alloys (where magnetic moments on Ni sites in alloys containing more than 33\% Cr are close to zero). Cr atoms prefer their magnetic moments ordered anti-ferromagnetically with respect to Fe and Ni moments. Their magnitudes are larger at low Cr concentration, and even at 25 \% Cr concentration they are fairly large (-2.44 $\mu_B$ and -2.53 $\mu_B$ for Fe$_2$CrNi and FeCrNi$_2$ structures). Because of strong anti-ferromagnetic interactions between Fe and Cr atoms, structures with large magnetic moments on Cr sites also have large magnetic moments on Fe sites (2.09 $\mu_B$ and 2.31 $\mu_B$ for Fe$_2$CrNi and FeCrNi$_2$ structures). An exception from this rule is the Fe-rich corner of the diagram, where fcc structures remain anti-ferromagnetic and the mean magnetic moment as well as average magnetic moments of the constituting components are equal or close to zero. \begin{figure*} \centering \begin{minipage}{.50\textwidth} \centering a)\includegraphics[width=.85\linewidth]{magmom_Fe_fcc} \end{minipage \begin{minipage}{.50\textwidth} \centering b)\includegraphics[width=.85\linewidth]{magmom_Fe_bcc} \end{minipage} \newline \begin{minipage}{.50\textwidth} \centering c)\includegraphics[width=.85\linewidth]{magmom_Cr_fcc} \end{minipage \begin{minipage}{.50\textwidth} \centering d)\includegraphics[width=.85\linewidth]{magmom_Cr_bcc} \end{minipage} \newline \begin{minipage}{.50\textwidth} \centering e)\includegraphics[width=.85\linewidth]{magmom_Ni_fcc} \end{minipage \begin{minipage}{.50\textwidth} \centering f)\includegraphics[width=.85\linewidth]{magmom_Ni_bcc} \end{minipage} \newline \caption{ (Color online) Magnetic moments (in $\mu_B$) of Fe (a, b), Cr (c, d) and Ni (e, f) on fcc (a, c, e) and bcc (b, d, f) lattices. } \label{fig:mag_FeCrNi_components} \end{figure*} Similarly to fcc structures, magnetic moments of Cr atoms on bcc lattice in the dilute Cr limit order anti-ferromagnetically with respect to those of Fe and Ni atoms, and their magnitudes decrease rapidly as a function of Cr content. For example, magnetic moments of Cr atoms in bcc Fe$_2$CrNi and FeCrNi$_2$ alloys are -0.19 $\mu_B$ and -0.12 $\mu_B$. In other words, they are an order of magnitude smaller than those found in fcc alloys. Furthermore, Cr-rich structures are not non-magnetic, as they are in the fcc case, but anti-ferromagnetic. Unlike Cr atoms, the magnitudes of magnetic moments on Fe and Ni sites are larger in bcc than in fcc alloys. Average moments on Fe sites are larger than 1 $\mu_B$ for most of the compositions, with the maximum value of 2.94 $\mu_B$ corresponding to FeCrNi$_{14}$ structure in the Ni-rich corner of the diagram. Average magnetic moments of Ni atoms are close to zero only in bcc Cr-rich Cr-Ni binary alloys. As the Fe content increases, the average magnetic moment of Ni atoms increases, too, reaching a maximum value of 0.86 $\mu_B$ in the Fe-rich corner, modelled by Fe$_{14}$CrNi structure. \section{Finite temperature phase stability of Fe-Cr-Ni alloys} \subsection{Enthalpy of formation} The finite temperature phase stability of Fe-Cr-Ni alloys was analyzed using quasi-canonical MC simulations and ECIs derived from DFT calculations. MC simulations were performed for 63 different compositions spanning all the binary and ternary Fe-Cr-Ni alloys on a 10\% composition mesh for each of the three constituents of the alloy, and additional 12 compositions with Cr and Ni content varying from 5\% to 35\% and from 25\% to 45\%, respectively, to increase the composition mesh density in the vicinity of the fcc-bcc phase transition line. Enthalpies of formation of fcc and bcc alloys at 300 K are shown in Fig. \ref{fig:Enthalpies_low_temp}. In fcc and bcc alloys there is a large region of concentrations where enthalpies of mixing are negative, coloured blue in Fig. \ref{fig:Enthalpies_low_temp}. Negative enthalpies of formation correspond to the fact that alloys decompose into mixtures of intermetallic phases. The negative formation enthalpies of fcc Fe-Cr-Ni alloys are mainly due to the formation of fcc FeNi, FeNi$_3$, FeNi$_8$ and CrNi$_2$ binary phases and fcc Fe$_2$CrNi ternary phase. In bcc Fe-Cr-Ni alloys, a negative enthalpy of formation is primarily due to the formation of Fe-Cr $\alpha$-phase and Fe$_4$Ni$_5$ VZn-like phase, where the latter is the most stable Fe-Ni phase on bcc lattice. The fact that Fe$_4$Ni$_5$ VZn-like phase is not observed experimentally is because it is significantly less stable than the corresponding fcc phase. \begin{figure*} \centering \begin{minipage}{.50\textwidth} \centering a)\includegraphics[width=.9\linewidth]{Form_fcc_300K} \end{minipage \begin{minipage}{.50\textwidth} \centering b)\includegraphics[width=.9\linewidth]{Form_bcc_300K} \end{minipage} \caption{ (Color online) Enthalpies of formation (in eV/atom) of fcc (a) and bcc (b) alloys calculated using MC simulations at 300K. } \label{fig:Enthalpies_low_temp} \end{figure*} \begin{figure*} \centering \begin{minipage}{.50\textwidth} \centering a)\includegraphics[width=.9\linewidth]{H_diff_300K} \end{minipage \begin{minipage}{.50\textwidth} \centering b)\includegraphics[width=.9\linewidth]{F_diff_300K} \end{minipage} \newline \begin{minipage}{.50\textwidth} \centering c)\includegraphics[width=.9\linewidth]{H_diff_600K} \end{minipage \begin{minipage}{.50\textwidth} \centering d)\includegraphics[width=.9\linewidth]{F_diff_600K} \end{minipage} \newline \begin{minipage}{.50\textwidth} \centering e)\includegraphics[width=.9\linewidth]{H_diff_900K} \end{minipage \begin{minipage}{.50\textwidth} \centering f)\includegraphics[width=.9\linewidth]{F_diff_900K} \end{minipage} \caption{ (Color online) Difference between enthalpies of formation (a,c) and free energies of formation (b,d) of fcc and bcc alloys, in eV/atom units, predicted using MC simulations for 300 K (a,b), 600 K (c,d) and 900 K (e,f). Black solid lines separate the Ni-rich region of stability of fcc alloys and the region of stability of bcc alloys predicted using the enthalpy and free energy criteria. } \label{fig:Energy_diff} \end{figure*} Having evaluated the difference between formation enthalpies of fcc and bcc alloys at 300K, we can now separate regions of stability of fcc and bcc alloys defined by the formation enthalpy criterion, see Fig. \ref{fig:Energy_diff}(a). The fcc-bcc phase transition lines determined using the same criterion at 600 K and 900 K, see Figs. \ref{fig:Energy_diff}(c) and \ref{fig:Energy_diff}(e), show that fcc and bcc phases remain stable broadly within the same composition ranges at 300 K and 600 K, whereas at 900 K we observe that the region of stability of fcc alloys shrinks in comparison with the region of stability of bcc alloys. Analyzing the stability of fcc and bcc alloys using their enthalpies of formation is convenient since one compares values derived from MC simulations with those computed directly by DFT for each representative alloy configuration. However, the enthalpy criterion of phase stability is valid only at relatively low temperatures. At high temperatures one should take into account the configurational entropy as well as vibrational and magnetic entropy contributions to the entropy and enthalpy of formation. In this study we do not treat vibrational entropy effects. The magnetic and configurational entropy contributions to the free energy of formation of fcc and bcc Fe-Cr-Ni alloys, and their effect on fcc-bcc phase stability, are analyzed in Section IV.D. We first discuss the enthalpies of formation, for which theoretical values can be validated by experimental data \cite{Kubaschewski1967,Dench1963}, and the magnetic contribution to the enthalpy of formation, which plays a significant part at high temperatures. Table \ref{tab:Fe} shows that pure iron at low temperatures is stable in bcc $\alpha$-phase whereas at 1185 K \cite{Chen2001} it transforms into the fcc $\gamma$-phase, and then back into the bcc $\delta$-phase. In order to investigate the formation enthalpies of alloys at high temperature one should at least take into account the effect of thermal magnetic excitations in Fe. The following correction can then be applied to the formation enthalpies of alloys in the high temperature limit. It is proportional to the concentration of Fe and is based on results given in Fig. 2 of Ref. \onlinecite{Lavrentiev2010}: \begin{equation} \Delta H_{lat}^{corr}\approx c_{Fe}\left[\left(E_{lat}(T)-E_{lat}(0)\right)-\left(E_{GS}(T)-E_{bcc}(0)\right)\right] , \label{eq:Correction_to_enthalpy} \end{equation} where $lat=fcc,bcc$, $E_{lat}(0)$ and $E_{lat}(T)$ are the energies of Fe on $lat$ at 0K and at temperature $T$ and $E_{GS}(T)$ is the temperature-dependent energy of the ground state, which is {\it either} fcc {\it or} bcc. The magnetic contribution to the enthalpy of formation of alloys described above is very important for predicting the position of the fcc-bcc phase transition line based on the formation enthalpy criterion. Figs. \ref{fig:Energy_diff_1600K}(a) and \ref{fig:Energy_diff_1600K}(c) show that the Ni-rich region of the composition diagram, where fcc phase has lower enthalpy of formation than bcc phase, is significantly larger and agrees better with the available experimental findings and CALPHAD simulations (see e.g. Fig. 7(a) in Ref. \onlinecite{Tomiska2004} with results at 1573 K). The enthalpies of formation of the most stable crystal structures of Fe-Cr-Ni alloys computed using MC simulations at 1600 K with magnetic correction applied, are compared to experimental data from Refs. \onlinecite{Kubaschewski1967,Dench1963} in Fig. \ref{fig:Enthalpy_triangle_1565K}, and the predictions agree with experiment very well. The enthalpy of formation treated as a function of temperature was examined by using MC simulations more extensively for one particular composition, Fe$_{70}$Cr$_{20}$Ni$_{10}$, close to the composition of austenitic 304 and 316 steels\cite{Piochaud2014}. As shown in Fig. \ref{fig:Enthalpies_low_temp}, at 300K, 600K and 900K the Fe$_{70}$Cr$_{20}$Ni$_{10}$ alloy belongs to the Fe-rich region of stability of bcc alloys, in agreement with experimental data and CALPHAD simulations (see e.g. Fig. 6 of Ref. \onlinecite{Franke2011} referring to 500 $^{\circ}$C). At 1600K, with the above magnetic correction applied, fcc alloy is more stable than bcc alloy, and its calculated enthalpy of formation of 0.030 eV/atom obtained from MC simulations at 1600K, is close to the experimental value of 0.035 eV/atom measured at 1565 K\cite{Kubaschewski1967}. Since austenitic stainless steels are formed by rapid cooling from approximately 1323K, we also analyze phase stability of Fe$_{70}$Cr$_{20}$Ni$_{10}$ alloy at 1300K \cite{Ferry2006,Majumdar1984}. Similarly to the 1600K case, after applying the magnetic correction, we find that fcc alloys have lower formation enthalpy than bcc alloys. The stability of various magnetic configurations of Fe$_{70}$Cr$_{20}$Ni$_{10}$ was analyzed using spin-polarized DFT calculations for the fcc structure with 256 atoms, derived from MC simulations at 1300K. As shown in Table \ref{tab:results_Fe70Cr20Ni10}, AFMSL and FM configurations are more stable than the AFMDL configuration, and the formation enthalpies of the two former ones are 0.018 eV/atom and 0.015 eV/atom higher than the value obtained from equilibrium MC simulations at 1300K. Results for the MC-generated structure are compared in Table \ref{tab:results_Fe70Cr20Ni10} also with enthalpies of formation of various magnetic configurations performed using fcc SQS with 256 atoms given in Ref. \onlinecite{Piochaud2014}. The energy of the most stable AFMSL configuration on SQS is 0.049 eV/atom higher than the energy of the most stable FM configuration realized on the MC-generated structure, and 0.064 eV/atom higher than the energy evaluated using equilibrium MC simulations at 1300K. Since the MC model simulations supplemented by magnetic correction were successfully validated against experimentally observed enthalpies of formation, as described above, we conclude that SQS-based calculations overestimate the formation enthalpy of the relevant alloy composition. Hence, even at high temperatures, configurations generated using CE combined with MC simulations describe Fe$_{70}$Cr$_{20}$Ni$_{10}$ alloy better than SQS. \begin{figure*} \centering \begin{minipage}{.50\textwidth} \centering a)\includegraphics[width=.9\linewidth]{H_diff_1600K_without} \end{minipage \begin{minipage}{.50\textwidth} \centering b)\includegraphics[width=.9\linewidth]{F_diff_1600K_without2} \end{minipage} \newline \begin{minipage}{.50\textwidth} \centering c)\includegraphics[width=.9\linewidth]{H_diff_1600K_with} \end{minipage \begin{minipage}{.50\textwidth} \centering d)\includegraphics[width=.9\linewidth]{F_diff_1600K_with2} \end{minipage} \caption{ (Color online) Difference between enthalpies of formation (a,c) and free energies of formation (b,d) of fcc and bcc alloys, in eV/atom units, calculated using MC simulations at 1600 K without (a,b) and with magnetic correction applied to the formation enthalpies (c) and free energies of formation (d). Black solid lines separate the Ni-rich region of stability of fcc alloys and the region of stability of bcc alloys predicted using the enthalpy and free energy criteria. } \label{fig:Energy_diff_1600K} \end{figure*} \begin{figure*} \begin{minipage}{.50\textwidth} \centering a)\includegraphics[width=.9\linewidth]{Form_1600K_with_corr} \end{minipage \begin{minipage}{.50\textwidth} \centering b)\includegraphics[width=.9\linewidth]{Form_1565K_Expt} \end{minipage} \caption{ (Color online) (a) Enthalpies of formation (in eV/atom) for the most stable crystal structures of Fe-Cr-Ni alloys computed using MC simulations at 1600K with magnetic correction applied, compared to experimental data (b) from Refs. \onlinecite{Kubaschewski1967} and \onlinecite{Dench1963}.} \label{fig:Enthalpy_triangle_1565K} \end{figure*} \begin{table} \caption{Enthalpies of formation of Fe$_{70}$Cr$_{20}$Ni$_{10}$ alloy derived from MC simulations at 1300K and 1600K, compared with experimental values measured at 1565 K\cite{Kubaschewski1967}, and with DFT energies computed for SQS and MC-generated atomic structures. \label{tab:results_Fe70Cr20Ni10}} \begin{ruledtabular} \begin{tabular}{cccc} & without corr. & with corr. & Expt. \\ \hline \multicolumn{3}{l}{MC at 1600 K} & \\ bcc & 0.028 & 0.045 & \\ fcc & 0.087 & 0.030 & 0.035 \\ \multicolumn{3}{l}{MC at 1300 K} & \\ bcc & 0.017 & 0.029 & \\ fcc & 0.051 & -0.006 & \\ \hline \multicolumn{3}{l}{DFT (SQS)$^a$} & \\ fcc AFMSL & 0.115 & & \\ fcc AFMDL & 0.126 & & \\ fcc FM & 0.116 & & \\ \hline \multicolumn{3}{l}{DFT (MC structure)$^b$} & \\ fcc AFMSL & 0.069 & & \\ fcc AFMDL & 0.105 & & \\ fcc FM & 0.066 & & \\ \end{tabular \end{ruledtabular} \begin{flushleft} $^a$ SQS structure from Ref. \onlinecite{Piochaud2014}.\\ $^b$ Structure generated using MC simulations performed at 1300 K. \\ \end{flushleft} \end{table} \subsection{Order-disorder transitions} \begin{figure*} \centering \begin{minipage}{.50\textwidth} \centering a)\includegraphics[width=.85\linewidth]{Order_fcc} \end{minipage \begin{minipage}{.50\textwidth} \centering b)\includegraphics[width=.85\linewidth]{Order_bcc} \end{minipage} \caption{(Color online) Order-disorder temperatures of fcc (a) and bcc (b) Fe-Cr-Ni alloys computed using Monte Carlo simulations. Order-disorder temperatures for pure elements are assumed to be 0 K. \label{fig:T_order_ternary}} \end{figure*} There is direct experimental evidence showing the presence of chemical order in Fe-Cr-Ni alloys. Bcc alloys at low temperatures segregate, with intermetallic Fe-Cr $\alpha$-phase representing the only known exception, whereas fcc alloys form austenitic steels exhibiting the formation of chemically ordered phases\cite{Dimitrov1986,Marwick1987,Cenedese1984,Menshikov1997}. \begin{figure} \includegraphics[width=\columnwidth]{SRO_Fe2CrNi} \caption{(Color online) Enthalpies of mixing and short-range order parameters (SRO) computed as functions of temperature for Fe-Cr, Cr-Ni and Fe-Ni atomic pairs in the first coordination shell of fcc Fe$_{50}$Cr$_{25}$Ni$_{25}$ alloy. \label{fig:SRO_Fe2CrNi}} \end{figure} Order-disorder phase transition temperatures correspond to inflection points on the energy versus temperature curves. Below a phase transition temperature, a chemically ordered phase is stable. For example, the Fe$_2$CrNi intermetallic phase remains ordered below 650K, whereas chemical order between Fe and Cr, and Cr and Ni pairs of atoms in fcc Fe$_{50}$Cr$_{25}$Ni$_{25}$ alloy vanishes above 1450 K, see Fig. \ref{fig:SRO_Fe2CrNi}. Ordering temperatures computed for all the fcc ground states of Fe-Cr-Ni alloys are given in Table \ref{tab:T_ord_structures}. Fcc Fe$_3$Cr L1$_2$ phase has the highest ordering temperature of 1550K. This phase is however less stable than the bcc phase with the same composition. Ordering temperatures predicted for FeNi, FeNi$_3$ and CrNi$_2$ alloys are in reasonable agreement with experimental data. The highest order-disorder transition temperatures $T_{ord-disord}$, above which alloys can be described as disordered, were found using Monte Carlo simulations for the entire range of alloy compositions, and are shown in Fig. \ref{fig:T_order_ternary}. $T_{ord-disord}$ of fcc Fe-Cr-Ni alloys shown in Fig. \ref{fig:T_order_ternary}(a) varies non-linearly as a function of composition, exhibiting even some local maxima. A local maximum near the FeNi$_3$ phase is in fact expected since FeNi$_3$ forms a L1$_2$ phase with relatively high ordering temperature. A local maximum around another experimentally known binary phase, CrNi$_2$ (MoPt$_2$), can also be recognized in Fig. \ref{fig:T_order_ternary}, however it is not as pronounced as in the FeNi$_3$ case. Two other maxima are less strongly pronounced. The first one corresponds to fcc alloys with Cr content between 25\% and 50\%, and is not very important for applications, since fcc alloys in this composition range are less stable than bcc alloys. More significant is the interval of compositions corresponding to Cr content from 10\% to 50\% and Ni content from 0\% to 50\%, which partially overlaps with the range of compositions of austenitic steels. Many Fe-Cr-Ni chemically ordered alloys are inside this composition range, with examples including Fe$_{64}$Cr$_{16}$Ni$_{20}$ and Fe$_{59}$Cr$_{16}$Ni$_{25}$ \cite{Dimitrov1986}, Fe$_{56}$Cr$_{21}$Ni$_{23}$ \cite{Cenedese1984}, and Fe$_{66.2}$Cr$_{17.5}$Ni$_{14.5}$Mo$_{2.8}$ \cite{Sharma1978}. All the bcc Fe-Cr-Ni alloys are predicted to have high order-disorder temperatures, see Fig. \ref{fig:T_order_ternary}(b). The alloys exhibit short range order even at temperatures close to melting. These predicted high order-disorder temperatures are likely to be over-estimated due to the fact that our CE-based Monte Carlo simulations neglect vibrational and magnetic contributions\cite{Walle2002a,Bonny2009}. \begin{table} \caption{Enthalpies of mixing and order-disorder transition temperatures predicted for several intermetallic phases of Fe-Cr-Ni alloys. \label{tab:T_ord_structures}} \begin{ruledtabular} \begin{tabular}{ccccc} Structure & $\Delta H_{mix}$ (eV) & $T_{ord}$ (K) & $T_{ord}^{Expt.}$ (K) \\ \hline fcc FeNi & -0.103 & 650 & 620\cite{Massalski1990} \\ fcc FeNi$_3$ & -0.116 & 950 & 790\cite{Massalski1990} \\ fcc FeNi$_8$ & -0.053 & 550 & \\ fcc Fe$_3$Ni$_2$ & -0.082 & 550 & \\ fcc CrNi$_2$ & -0.155 & 750 & 863\cite{Massalski1990} \\ fcc Cr$_2$Ni & -0.182 & 1250 & \\ fcc Cr$_3$Ni & -0.153 & 1150 & \\ fcc Fe$_3$Cr & -0.103 & 1550 & \\ fcc FeCr$_2$ & -0.119 & 850 & \\ fcc FeCr$_8$ & -0.052 & 350 & \\ fcc Fe$_2$CrNi & -0.164 & 650 & \\ \end{tabular \end{ruledtabular} \end{table} Analysis showing how the predicted $T_{ord-disord}$ varies as a function of alloy composition confirms the experimentally observed reduction of chemical ordering in (FeNi$_3$)$_{1-x}$Cr$_x$ alloys annealed at 486 $^{\circ}$C ($=$759K) as a function of Cr content in the composition interval from $x$=0.0 to 0.17\cite{Marwick1987}. The values of $T_{ord-disord}$ obtained from MC simulations for FeNi$_3$ and (FeNi$_3$)$_{0.8}$Cr$_{0.2}$ alloys are 950K and 750K, respectively. Alloys with lower Cr content have order-disorder transition temperatures significantly higher than the annealing temperature used in the above experiments \cite{Marwick1987}. Values of $T_{ord-disord}$ in alloys with high Cr content are lower than the above annealing temperature. \subsection{Short-range order parameters} Chemical order in alloys is characterized by the Warren-Cowley short-range order parameters, $\alpha_1^{i-j}$ and $\alpha_2^{i-j}$, for the first (1NN) and second (2NN) nearest neighbour coordination shells. These parameters are calculated from Eq. \ref{eq:SRO_pairs} using correlation functions deduced from MC simulations. MC simulations were performed, assuming various temperatures, for several binary and four ternary alloy compositions: Fe$_{56}$Cr$_{21}$Ni$_{23}$, Fe$_{42.5}$Cr$_{7.5}$Ni$_{50}$, Fe$_{38}$Cr$_{14}$Ni$_{48}$ and Fe$_{34}$Cr$_{20}$Ni$_{46}$, for which experimental SRO parameters were published in Refs. \onlinecite{Cenedese1984} and \onlinecite{Menshikov1997}. SRO parameters computed for binary alloys agree with experimental data, see Table \ref{tab:SRO_binary_alloys}. Comparison with experimental data for ternary Fe-Cr-Ni alloys is given in Fig. \ref{fig:SRO_Cenedese&Menshikov} and in Table \ref{tab:SRO_Cenedese&Menshikov}. MC simulations performed for fcc Fe$_{56}$Cr$_{21}$Ni$_{23}$ alloy \cite{Cenedese1984} show that it is characterized by pronounced Cr-Ni ordering, whereas at the same time there is no Fe-Ni ordering. Values of $\alpha_1^{Fe-Ni}$ and $\alpha_1^{Cr-Ni}$ are in excellent agreement with experimental observations. The calculated SRO parameter for Fe and Cr atoms is negative, in agreement with experimental observations, although the magnitude of this parameter predicted by calculations is larger. This may again be due to the fact that vibrational and magnetic contributions were neglected \cite{Walle2002a}. The effect of lattice vibrations on ordering in the bcc Fe-Cr system was noted in Refs. \onlinecite{Lavrentiev2007,Lavrentiev2010,Bonny2009}. The effect of Cr on SRO in Fe-Cr-Ni alloys was analyzed, using MC simulations, for three compositions Fe$_{42.5}$Cr$_{7.5}$Ni$_{50}$, Fe$_{38}$Cr$_{14}$Ni$_{48}$ and Fe$_{34}$Cr$_{20}$Ni$_{46}$, which are the compositions investigated experimentally in Ref. \onlinecite{Menshikov1997}. As expected, the absolute values of SRO parameters increase with decreasing temperature for both 1NN and 2NN (see Table \ref{tab:SRO_Cenedese&Menshikov}). All the 2NN SRO parameters are positive for these three alloys. $\alpha_1^{Fe-Ni}$ and $\alpha_1^{Fe-Cr}$ are negative and their absolute values decrease as functions of Cr content. An interesting result is that the sign of $\alpha_1^{Cr-Ni}$ changes from positive for Fe$_{42.5}$Cr$_{7.5}$Ni$_{50}$ to negative for Fe$_{34}$Cr$_{20}$Ni$_{46}$ alloy. Fe$_{38}$Cr$_{14}$Ni$_{48}$ alloy with intermediate Cr content has positive $\alpha_1^{(Fe-Cr)}$ only at relatively low temperatures close to 600K. We compare our theoretical predictions with measured SRO parameters involving Ni atoms and 'average' (Fe,Cr) atoms in Fe$_{42.5}$Cr$_{7.5}$Ni$_{50}$, Fe$_{38}$Cr$_{14}$Ni$_{48}$ and Fe$_{34}$Cr$_{20}$Ni$_{46}$ alloys, quenched rapidly from 1323K, annealed at 873K and irradiated at 583K with 2.5 MeV electrons\cite{Menshikov1997}. The authors of Ref. \onlinecite{Menshikov1997} neglected ordering between Fe and Cr atoms, arguing that there was no evidence for the occurrence of stable Fe-Cr compounds at low temperatures. Their assumption was based also on experimental observations by Cenedese {\it et al.}\cite{Cenedese1984} who found that in fcc Fe$_{56}$Cr$_{21}$Ni$_{23}$ alloy only Cr and Ni atoms were ordered. To compare results of MC simulations with experimentally measured \cite{Menshikov1997} SRO parameters involving Ni and 'average' (Fe,Cr) atoms, we treat ternary Fe-Cr-Ni alloys as a pseudo-binary alloy of composition Ni$_x$(FeCr)$_{1-x}$. We define an effective SRO parameter involving Ni and (Fe,Cr) atoms as \begin{equation} \alpha_n^{(Fe,Cr)-Ni} = \frac{c_{Fe}}{c_{Fe}+c_{Cr}}\alpha_n^{Fe-Ni} + \frac{c_{Cr}}{c_{Fe}+c_{Cr}}\alpha_n^{Cr-Ni}. \label{eq:alpha_FeCr_Ni} \end{equation} Values of $\alpha_1^{(Fe,Cr)-Ni}$ defined in this way and calculated using MC simulations for 1300K are in excellent agreement with experimental observations for alloy samples quenched rapidly from 1323K. Despite the fact that experimental measurements for samples irradiated at 583K cannot be directly compared with MC simulations at 600K, experimental observations showing more pronounced chemical order in Fe$_{34}$Cr$_{20}$Ni$_{46}$ sample irradiated at 583K in comparison with Fe$_{38}$Cr$_{14}$Ni$_{48}$ sample are in agreement with our predictions. The occurrence of chemical order in Fe-Cr-Ni alloys can be explained by analysing interactions between pairs of Fe-Cr, Fe-Ni and Cr-Ni atoms, $V_n^{ij}$. They were derived from two-body effective cluster interaction parameters for fcc and bcc Fe-Cr-Ni alloys listed in Table \ref{tab:ECI_def_ternary}. Assuming that many-body interactions are small, we find that $V_n^{ij}$ are related to $J_{2,n}^{(s)}$ and can be calculated using Eq. \ref{eq:VvsJ_Matrix}. $V_n^{Fe-Ni}$, $V_n^{Fe-Cr}$ and $V_n^{Cr-Ni}$ computed for fcc and bcc ternary alloys are compared with values derived for binary alloys in Fig. \ref{fig:ECI_Vij}. All the 1NN chemical pairwise interactions computed for bcc ternary alloys are even more negative than those computed for binary alloys. This means that repulsion between Fe-Cr, Fe-Ni and Cr-Ni atoms in ternary alloys is even stronger than in binary alloys. On the other hand, in ternary fcc alloys the 1NN chemical pairwise interaction between Fe and Ni atoms vanishes almost completely. This explains why the SRO parameter involving Fe and Ni atoms in the first nearest neighbour coordination shell measured in Ref. \onlinecite{Cenedese1984} nearly vanishes. This also explains the observed decrease of atomic ordering in FeNi$_3$ alloys following the addition of Cr \cite{Marwick1987}. Large 1NN effective pairwise Fe-Cr and Cr-Ni interactions ($V_1^{Fe-Cr}$ and $V_1^{Cr-Ni}$ are correspondingly smaller and larger in ternary alloys in comparison with binary alloys), and the relatively large 2NN effective interactions between these atoms also explain the pronounced atomic ordering in the majority of fcc Fe-Cr-Ni alloys. The sign pattern of the first three nearest-neighbour chemical pairwise interactions in all the binary and ternary fcc alloys remains the same. The first and third nearest-neighbour (3NN) pair interactions are positive and the second nearest neighbour interaction is negative. This favours the unlike atoms occupying the 1NN and 3NN coordination shells, and the like atoms occupying the 2NN shell, see Eq.\ref{eq:CE_vs_V}. Such a pattern of signs of NN interactions favours not only intermetallic L1$_2$ and MoPt$_2$-like phases, which occur in fcc binary alloys, but also the L1$_2$-based (Cu$_2$ZnNi-like) Fe$_2$CrNi ternary phase, which is the global ground state of ternary Fe-Cr-Ni alloys. \begin{table*} \caption{Short-range order parameters for selected binary alloys calculated using Monte Carlo simulations at various temperatures $T$, compared with experimental data. \label{tab:SRO_binary_alloys}} \begin{ruledtabular} \begin{tabular}{cccccccc} &$T$ (K) & MC (this study) & MC (Others) & Expt. & MC & MC \cite{Rahaman2014} & Expt. \\ \hline fcc alloys & & \multicolumn{3}{c}{$\alpha_1$} & \multicolumn{3}{c}{$\alpha_2$} \\ Fe$_{25}$Ni$_{75}$ & 1300 & -0.096 & & -0.099$^c$ & 0.097 & & 0.116$^c$ \\ Fe$_{30}$Ni$_{70}$ & & -0.102 & & -0.088$^c$ & 0.105 & & 0.049$^c$ \\ Fe$_{50}$Ni$_{50}$ & & -0.071 & & -0.073$^c$ & 0.082 & & 0.042$^c$ \\ Fe$_{60}$Ni$_{40}$ & & -0.043 & & -0.058$^c$ & 0.058 & & 0.089$^c$ \\ Fe$_{65}$Ni$_{35}$ & & -0.018 & & -0.051$^c$ & 0.049 & & 0.034$^c$ \\ Fe$_{70}$Ni$_{30}$ & & -0.002 & & -0.033$^c$ & 0.031 & & 0.005$^c$ \\ Fe$_{65}$Ni$_{35}$ & 1100 & -0.022 & & -0.058$^d$ & 0.076 & & 0.052$^d$ \\ Cr$_{33}$Ni$_{67}$ & 1100 & -0.036 & -0.115$^a$ & -0.08$^e$ & 0.042 & 0.12\cite{Rahaman2014} & 0.05$^e$ \\ Cr$_{25}$Ni$_{75}$ & 1000 & -0.047 & -0.105$^a$ & -0.07$^f$ & 0.051 & 0.10\cite{Rahaman2014} & 0.045$^f$ \\ Cr$_{20}$Ni$_{80}$ & 800 & -0.029 & -0.125$^a$ & -0.10$^g$ & 0.057 & 0.115\cite{Rahaman2014} & 0.085$^g$ \\ \hline bcc alloys & & \multicolumn{3}{c}{$\alpha_{1+2}$} & & & \\ Fe$_{95}$Cr$_{5}$ & 700 & -0.044 & -0.049$^b$ & -0.05$^h$ & & & \\ Fe$_{93.75}$Cr$_{6.25}$ & & -0.056 & & & & & \\ Fe$_{90}$Cr$_{10}$ & & -0.071 & -0.080$^b$ & 0.00$^h$ & & & \\ Fe$_{85}$Cr$_{15}$ & & 0.138 & 0.309$^b$ & 0.065$^h$ & & & \\ \end{tabular \end{ruledtabular} \begin{flushleft} $^a$ Ref. \onlinecite{Rahaman2014} MC simulations. \\ $^b$ Ref. \onlinecite{Lavrentiev2007} MC simulations. \\ $^c$ Refs. \onlinecite{Gomankov1971,Menshikov1972} annealed at 1273 K. \\ $^d$ Ref. \onlinecite{Robertson1999} annealed at 1026 K. \\ $^e$ Ref. \onlinecite{Caudron1992} annealed at 1073 K. \\ $^f$ Ref. \onlinecite{Caudron1992} annealed at 993 K. \\ $^g$ Ref. \onlinecite{Schonfeld1988} annealed at 828 K. \\ $^h$ Ref. \onlinecite{Mirebeau1984} annealed at 703 K. \\ \end{flushleft} \end{table*} \begin{table} \caption{Short-range order parameters for Fe-Cr, Fe-Ni and Cr-Ni pairs in ternary alloys calculated using Monte Carlo simulations, and compared with experimental observations. (Fe,Cr)-Ni means average SRO involving Ni and average (Fe,Cr) atoms as defined by Eq. \ref{eq:alpha_FeCr_Ni}. \label{tab:SRO_Cenedese&Menshikov}} \begin{ruledtabular} \begin{tabular}{cccccc} & & \multicolumn{2}{c}{$\alpha_1$} & \multicolumn{2}{c}{$\alpha_2$} \\ & & MC & Expt. & MC & Expt. \\ \hline Fe$_{56}$Cr$_{21}$Ni$_{23}$ & & & & & \\ Fe-Ni & 1300 & 0.003 & 0.017$^a$ & -0.094 & -0.002$^a$ \\ Fe-Cr & & -0.280 & -0.009$^a$ & 0.781 & 0.043$^a$ \\ Cr-Ni & & -0.134 & -0.113$^a$ & 0.600 & 0.148$^a$ \\ Fe$_{42.5}$Cr$_{7.5}$Ni$_{50}$ & & & & & \\ Fe-Ni & & -0.069 & & 0.073 & \\ Fe-Cr & 1300 & -0.080 & & 0.087 & \\ Cr-Ni & & 0.015 & & 0.085 & \\ (Fe,Cr)-Ni & & -0.057 & -0.049$^b$ & 0.075 & 0.015$^b$ \\ Fe-Ni & & -0.099 & & 0.144 & \\ Fe-Cr & 900 & -0.158 & & 0.281 & \\ Cr-Ni & & 0.039 & & 0.134 & \\ (Fe,Cr)-Ni & & -0.077 & -0.093$^c$ & 0.142 & 0.134$^c$ \\ Fe-Ni & & -0.213 & & 0.681 & \\ Fe-Cr & 600 & -0.403 & & 0.975 & \\ Cr-Ni & & 0.180 & & 0.680 & \\ (Fe,Cr)-Ni & & -0.150 & -0.121$^d$ & 0.681 & 0.148$^d$ \\ Fe$_{38}$Cr$_{14}$Ni$_{48}$ & & & & & \\ Fe-Ni & & -0.054 & & 0.054 & \\ Fe-Cr & 1300 & -0.076 & & 0.141 & \\ Cr-Ni & & -0.014 & & 0.115 & \\ (Fe,Cr)-Ni & & -0.043$^b$ & -0.048 & 0.070 & 0.018$^b$ \\ Fe-Ni & & -0.076 & & 0.116 & \\ Fe-Cr & 900 & -0.276 & & 0.701 & \\ Cr-Ni & & -0.023 & & 0.382 & \\ (Fe,Cr)-Ni & & -0.062 & -0.091$^c$ & 0.188 & 0.082$^c$ \\ Fe-Ni & & -0.149 & & 0.648 & \\ Fe-Cr & 600 & -0.489 & & 0.983 & \\ Cr-Ni & & 0.097 & & 0.814 & \\ (Fe,Cr)-Ni & & -0.082 & -0.126$^d$ & 0.693 & 0.089$^d$ \\ Fe$_{34}$Cr$_{20}$Ni$_{46}$ & & & & & \\ Fe-Ni & & -0.035 & & 0.033 & \\ Fe-Cr & 1300 & -0.060 & & 0.168 & \\ Cr-Ni & & -0.031 & & 0.136 & \\ (Fe,Cr)-Ni & & -0.033 & -0.042$^b$ & 0.071 & 0.017$^b$ \\ Fe-Ni & & -0.035 & & 0.024 & \\ Fe-Cr & 900 & -0.300 & & 0.838 & \\ Cr-Ni & & -0.146 & & 0.563 & \\ (Fe,Cr)-Ni & & -0.077 & -0.088$^c$ & 0.224 & 0.086$^c$ \\ Fe-Ni & & -0.138 & & 0.538 & \\ Fe-Cr & 600 & -0.401 & & 0.994 & \\ Cr-Ni & & -0.133 & & 0.799 & \\ (Fe,Cr)-Ni & & -0.137 & -0.152$^d$ & 0.634 & 0.098$^d$ \\ \end{tabular \begin{flushleft} $^a$ Ref.\onlinecite{Cenedese1984} annealed at 1273 K. \\ $^b$ Ref.\onlinecite{Menshikov1997} quenched from 1323 K. \\ $^c$ Ref.\onlinecite{Menshikov1997} annealed at 873 K. \\ $^d$ Ref.\onlinecite{Menshikov1997} irradiated at 583 K with 2.5 MeV electrons. \\ \end{flushleft} \end{ruledtabular} \end{table} \begin{figure*} \centering \begin{minipage}{.50\textwidth} \centering a)\includegraphics[width=.85\linewidth]{SRO_Cenedese} \end{minipage \begin{minipage}{.50\textwidth} \centering b)\includegraphics[width=.85\linewidth]{SRO_Menshikov} \end{minipage} \caption{(Color online) (a) Short-range order parameters as functions of temperature, calculated for Fe-Cr, Cr-Ni and Fe-Ni pairs occupying two coordination shells in Fe$_{0.56}$Cr$_{0.21}$Ni$_{0.23}$ alloy, compared with experimental values from Ref. \onlinecite{Cenedese1984}; (b) 1NN SRO between Ni and average (Fe,Cr) atoms in Fe$_{42.5}$Cr$_{7.5}$Ni$_{50}$, Fe$_{38}$Cr$_{14}$Ni$_{48}$ and Fe$_{34}$Cr$_{20}$Ni$_{46}$ calculated at 600K, 900K and 1300K, and compared with experimental data taken from Ref. \onlinecite{Menshikov1997} and presented as a function of Cr content; (Fe,Cr)-Ni indicates average SRO between Ni and average (Fe,Cr) atoms obtained from Eq. \ref{eq:alpha_FeCr_Ni}. \label{fig:SRO_Cenedese&Menshikov}} \end{figure*} \begin{figure*} \centering \begin{minipage}{.50\textwidth} \centering a)\includegraphics[width=.85\linewidth]{ECI_FeCr_fcc_Vij} \end{minipage \begin{minipage}{.50\textwidth} \centering b)\includegraphics[width=.85\linewidth]{ECI_FeCr_bcc_Vij} \end{minipage} \newline \begin{minipage}{.50\textwidth} \centering c)\includegraphics[width=.85\linewidth]{ECI_FeNi_fcc_Vij} \end{minipage \begin{minipage}{.50\textwidth} \centering d)\includegraphics[width=.85\linewidth]{ECI_FeNi_bcc_Vij} \end{minipage} \newline \begin{minipage}{.50\textwidth} \centering e)\includegraphics[width=.85\linewidth]{ECI_CrNi_fcc_Vij} \end{minipage \begin{minipage}{.50\textwidth} \centering f)\includegraphics[width=.85\linewidth]{ECI_CrNi_bcc_Vij} \end{minipage} \newline \caption{ Effective interactions between different pairs of atoms: Fe-Cr (a,b), Fe-Ni (c,d) and Cr-Ni (e,f) on fcc (a,c,e) and bcc (b,d,f) lattices in ternary Fe-Cr-Ni and binary alloys. } \label{fig:ECI_Vij} \end{figure*} \subsection{Configurational entropy and free energy of formation} While by comparing enthalpies of formations we assess the low temperature phase stability of alloys, the investigation of high temperature phase stability requires comparing formation free energies of fcc and bcc phases. Evaluating the free energy requires computing both the enthalpy of formation and the configurational entropy of the alloy. Configurational entropy is defined as \begin{equation} S_{conf}(T)=\int_0^T\frac{C_{conf}(T')}{T'}dT', \label{eq:ConfEntropy} \end{equation} where the configurational contribution to the specific heat $C_{conf}$ is related to fluctuations of enthalpy of mixing at a given temperature \cite{Newman1999,Lavrentiev2009} through \begin{equation} C_{conf}(T)=\frac{\left\langle H_{mix}(T)^2\right\rangle-\left\langle H_{mix}(T)\right\rangle^2}{T^2}, \label{eq:ConfHeatCapacity} \end{equation} where $\left\langle H_{mix}(T)\right\rangle$ and $\left\langle H_{mix}(T)^2\right\rangle$ are the mean and mean square average enthalpies of mixing, respectively, computed by averaging over all the MC steps at the accumulation stage for a given temperature. The accuracy of evaluation of configurational entropy depends on temperature integration step in Eq. \ref{eq:ConfEntropy} and the number of MC steps performed at the accumulation stage. Test simulations showed that choosing a sufficiently small temperature integration step is particularly significant. Calculations of configurational entropy for all the alloy compositions below were performed with 2000 MC steps per atom at the thermalization and accumulation stages and with temperature step of $\Delta T = 10$ K. \begin{figure*} \centering \begin{minipage}{.50\textwidth} \centering a)\includegraphics[width=.85\linewidth]{Entropy_Fe2CrNi_v2} \end{minipage \begin{minipage}{.50\textwidth} \centering b)\includegraphics[width=.85\linewidth]{FreeEnergy_Fe2CrNi_v2} \end{minipage} \caption{(Color online) (a) Configurational specific heat $C_{conf}$, configurational entropy $S_{conf}$, (b) the enthalpy of formation $H_{form}$, the product of temperature and configurational entropy $-TS_{conf}$ and the free energy of formation $F_{form}$ of fcc and bcc Fe$_2$CrNi alloys. The black dotted line is the entropy of ideal random solid solution of Fe$_2$CrNi equal to 1.04$k_B$. \label{fig:Entropy_FreeEn_Fe2CrNi}} \end{figure*} Configurational specific heat and configurational entropy of fcc and bcc Fe$_2$CrNi alloys, treated as functions of temperature, are shown in Fig. \ref{fig:Entropy_FreeEn_Fe2CrNi}(a). Configurational specific heats of both alloys exhibit sharp peaks in the vicinity of order-disorder phase transition temperatures. For example, the first peak at 700 K in the specific heat curve of fcc Fe$_2$CrNi alloy refers to the temperature of ordering of Fe and Ni atoms, whereas the second peak at 1500 K refers to the ordering temperature of Fe-Cr and Cr-Ni atoms, see Fig. \ref{fig:SRO_Fe2CrNi}. Configurational entropy of fcc Fe$_2$CrNi alloy in the high temperature limit approaches the configurational entropy of ideal random solid solution for this composition given by the formula \begin{equation} S_{random}(T)=-k_B\sum_ic_iln(c_i). \label{eq:Conf_Entropy} \end{equation} Substituting atomic concentrations in this equation, we find that for Fe$_2$CrNi alloy the configurational entropy in the high temperature limit is equal to 1.04 $k_B$. Configurational entropy of bcc Fe$_2$CrNi alloy at 2000 K is lower. This is due to the fact that bcc alloy at 2000 K is still not fully random. For temperatures above the temperature of ordering of Fe and Ni atoms in fcc Fe$_2$CrNi alloy, the entropy of fcc alloy is always higher than that of bcc alloy. Hence, despite the fact that the enthalpy of formation of bcc alloy at high temperatures is lower than that of fcc alloy, the latter is always more stable according to the formation free energy criterion, see Fig. \ref{fig:Entropy_FreeEn_Fe2CrNi}(b). \begin{figure*} \centering \begin{minipage}{.50\textwidth} \centering a)\includegraphics[width=.9\linewidth]{Form_fcc_900K} \end{minipage \begin{minipage}{.50\textwidth} \centering b)\includegraphics[width=.9\linewidth]{Form_bcc_900K} \end{minipage} \newline \begin{minipage}{.50\textwidth} \centering c)\includegraphics[width=.9\linewidth]{TS_fcc_900K} \end{minipage \begin{minipage}{.50\textwidth} \centering d)\includegraphics[width=.9\linewidth]{TS_bcc_900K} \end{minipage} \newline \begin{minipage}{.50\textwidth} \centering e)\includegraphics[width=.9\linewidth]{Free_energy_fcc_900K} \end{minipage \begin{minipage}{.50\textwidth} \centering f)\includegraphics[width=.9\linewidth]{Free_energy_bcc_900K} \end{minipage} \caption{ (Color online) Enthalpies of formation (in eV/atom) (a,b), the product of temperature and configurational entropies $TS_{conf}$ (in eV/atom units) (c,d) and free energies of formation (in eV/atom units) (e,f) at 900 K computed using MC simulations for Fe-Cr-Ni alloys on fcc (a,c,e) and bcc (b,d,f) lattices. } \label{fig:EntropyFreeEn_900K} \end{figure*} Formation free energies of fcc and bcc Fe-Cr-Ni alloys at 900 K computed using MC simulations for the entire range of alloy compositions are shown in Figs. \ref{fig:EntropyFreeEn_900K}(e) and \ref{fig:EntropyFreeEn_900K}(f), together with their formation enthalpies and configurational entropies, see Figs. \ref{fig:EntropyFreeEn_900K}(a,b) and Figs. \ref{fig:EntropyFreeEn_900K}(c,d). Configurational entropy of fcc alloys is higher for most of the alloy compositions in comparison with that of bcc alloys. Hence the configurational entropy contribution to the formation free energies is more significant for fcc alloys than for bcc alloys. The region of stability of fcc alloys at 900 K defined using the free energy criterion is broader than the region of stability defined using the formation enthalpy criterion, see Figs. \ref{fig:Energy_diff}(e) and \ref{fig:Energy_diff}(f). At low temperatures the difference between fcc-bcc phase transition lines obtained using both criteria is negligible, see Figs. \ref{fig:Energy_diff}(a-d), whereas at high temperatures the role played by the configurational entropy effects is more pronounced, see Figs. \ref{fig:Energy_diff}(e-f) and \ref{fig:Energy_diff_1600K}(a-b). \section{Finite temperature magnetic properties of Fe-Cr-Ni alloys} Using the DFT database and Magnetic Cluster Expansion, we now investigate how magnetic properties of Fe-Cr-Ni ternary alloys vary as functions of temperature. Monte Carlo MCE simulations were performed using a 16384 atom simulation cell (16$\times$16$\times$16 fcc unit cells). At each MC step, a trial random variation of the magnetic moment of a randomly chosen atom is attempted and accepted or rejected according to the Metropolis criterion. Angular and longitudinal fluctuations of magnetic moments are relatively small, resulting in the formation of non-collinear magnetic configurations. We do not consider MC moves that change the sign of magnetic moment. The gap-less spectrum of magnetic excitations (magnons) in our model is described by small tilts of magnetic moments away from their equilibrium orientations. Both the thermalization and accumulation stages include on average 40 000 MC steps per atom. As an example of application of MCE to modelling low-temperature magnetic properties of a ternary alloy, we investigate how the total magnetic moment of a disordered (Fe$_{0.5}$Ni$_{0.5}$)$_{1-x}$Cr$_{x}$ alloy varies as a function of Cr content. We noted in Section III.E that the average magnetic moment of the alloy decreased rapidly as a function of chromium concentration. In MCE Monte Carlo simulations, ordered Fe-Ni alloy with L1$_0$ structure was chosen as the initial alloy configuration. The magnetic moment per atom in this structure was found to be 1.61 $\mu_B$, close to the DFT value of 1.63 $\mu_B$. Chromium content was then varied by replacing equal numbers of Fe and Ni atoms in their sublattices with Cr atoms, with positions of chromium atoms chosen at random. Figure \ref{fig:MCE_FeNi-Cr} shows the predicted variation of magnetic moment in the resulting alloy at low temperatures. With increasing Cr content, magnetization rapidly decreases, resulting in a completely non-magnetic system at $x_{Cr}$=0.4, in agreement with {\it ab initio} results of Section III.E, also illustrated in Figure \ref{fig:MCE_FeNi-Cr}. \begin{figure} \includegraphics[width=\columnwidth]{MCE_FeNi-Cr} \caption{(Color online) Total magnetic moment per atom in (Fe$_{0.5}$Ni$_{0.5}$)$_{1-x}$Cr$_{x}$ alloy as a function of chromium content $x$, predicted by MCE. DFT results (Sec. III.E) are shown by red circles. \label{fig:MCE_FeNi-Cr}} \end{figure} While random alloys with composition close to Fe$_{50}$Cr$_{25}$Ni$_{25}$ are almost entirely anti-ferromagnetic, ordered alloys with the same composition have non-vanishing total magnetic moment. At low temperature, magnetic moments are collinear. The Cr moments are anti-ferromagnetically ordered with respect to Fe moments and have almost the same magnitude, while the magnetic moments of Ni atoms are smaller and ordered ferromagnetically with respect to the Fe moments. Finite temperature magnetic properties were investigated using MC simulations performed using large simulation cells. Magnetic moments of each of the three components of ordered Fe$_2$CrNi alloy, treated as functions of temperature, are shown in Figure \ref{fig:MCE_Fe2CrNi}. Their values at low temperature are in reasonable agreement with DFT, for example magnetic moments of Fe, Cr and Ni obtained from MCE simulations are 2.7, -2.2 and 0.37 $\mu_B$, whereas DFT predictions are 2.08, -2.44 and 0.15 $\mu_B$, respectively, see Table \ref{tab:Enthalpies_of_GS}. The alloy remains magnetic at fairly high temperatures close to 850-900 K. The effect is similar to that found in fcc Fe-Ni, where chemically ordered FeNi$_3$ alloy has higher Curie temperature than pure Ni. In ternary Fe$_2$CrNi alloy, the ferromagnetically ordered structure of the alloy owes its stability to strong anti-ferromagnetic interactions between (Fe, Ni) and Cr atoms. \begin{figure} \includegraphics[width=\columnwidth]{MCE_Fe2CrNi} \caption{(Color online) Temperature dependence of the total magnetic moment of ordered Fe$_2$CrNi alloy, and magnetic moments of atoms forming the alloy. \label{fig:MCE_Fe2CrNi}} \end{figure} Another important application of the Magnetic Cluster Expansion to the high-temperature properties of the Fe-Cr-Ni system is the study of the relative stability of fcc and bcc structures. Previously \cite{Lavrentiev2010} we have used the MCE in order to find the free energy difference between fcc and bcc Fe in the whole range of temperatures from 0 K to the melting point. Here, we use these results in order to estimate the magnetic correction to the free energy of alloy formation. For this correction we used formulae analogous to Eq. \ref{eq:Correction_to_enthalpy} with the free energy difference between fcc and bcc Fe taken from Fig. 2 of Ref. \onlinecite{Lavrentiev2010}. Fig. \ref{fig:Energy_diff_1600K} shows that both magnetic and configurational effects are important. However, it is also apparent that the magnetic correction to free energy difference, compare Figs. \ref{fig:Energy_diff_1600K}(b) and \ref{fig:Energy_diff_1600K}(d), is less pronounced than that applied to the enthalpy difference, see Figs. \ref{fig:Energy_diff_1600K}(a) and \ref{fig:Energy_diff_1600K}(c). This is related to the fact that even though the fcc phase of Fe is strongly stabilized at high temperatures in terms of enthalpy, the free energy of bcc Fe with magnetic effects taken into account is still lower than that of fcc Fe, see Fig. 2 of Ref. \onlinecite{Lavrentiev2010}. As discussed in Ref. \onlinecite{Lavrentiev2010}, only after the inclusion of vibrational effects can the stability of fcc phase at high temperatures be predicted correctly. It can be deduced that the vibrational contribution to free energy is also important at high temperatures for Fe-Cr-Ni alloys. The corresponding study will be performed in our future research. Concluding this section, we note that although the parametrization of MCE Hamiltonian involved a number of approximations, MCE predictions are in good agreement with the low temperature DFT data. In the high temperature limit, MC simulations show that interplay between chemical and magnetic degrees of freedom gives rise to the high Curie temperature of ordered Fe$_2$CrNi alloy. Further improvement in the accuracy of the MCE model is expected to provide a means for investigating temperature-dependent magnetic and configurational order over the entire range of alloy compositions. \section{Conclusions} We have investigated the stability of the fcc and bcc phases of ternary magnetic Fe-Cr-Ni alloys, using a combination of first-principles DFT calculations and Monte Carlo simulations, involving both conventional (CE) and magnetic (MCE) cluster expansion approaches. Detailed derivation of a general expression for the CE enthalpy of mixing for a ternary alloy system is presented, where average cluster functions are defined as products of orthogonal point functions. An explicit analytical relationship between chemical SRO and effective pair-wise interactions, involving different atomic species, is established and applied to the analysis of SRO in Fe-Cr-Ni alloys and the interpretation of experimental data. Using a DFT database of 248 fcc and 246 bcc structures, we assessed fcc and bcc phase stability of this ternary alloy system. Effective cluster interaction parameters for fcc and bcc binaries and ternaries have been derived and cross-validated against DFT data. Strong deviations from Vegard's law for atomic volumes treated as functions of alloy composition stem from magnetic interactions. The predicted average total and local magnetic moments treated as functions of Ni concentration in the ground-state bcc and fcc structures of Fe-Ni alloys are in good agreement with experimental data. Calculations have not only helped identify ground-state structures of the three binary alloys, but also predicted the fcc-like Fe$_{2}$CrNi ternary compound as the most stable ground-state ternary intermetallic system with negative enthalpy of formation of -0.164 eV/atom and the lowest order-disoder transition temperature of 650 K. Both DFT and MCE simulations show that the phase stability of the Fe$_{2}$CrNi structure is primarily determined by strong anti-ferromagnetic interactions between Fe and Ni atoms with Cr atoms. Analysis of the relative phase stability of ternary fcc and bcc phases at various temperatures in terms of formation enthalpies and formation free energies shows that configurational entropy plays an important part at high temperature in stabilizing fcc alloys with respect to bcc alloys. Preliminary incorporation of magnetic entropy for free energy differences in the Fe-rich corner shows that non-collinear magnetic effects are important at high temperatures. Excellent agreement between calculations and experimental data on enthalpies of formation at 1600 K also shows that magnetic contribution plays a significant part, correcting the deficiencies of conventional CE treatment of Fe-Cr-Ni alloys. We have calculated the Warren-Cowley short range order parameters at various temperatures and found good agreement with experimental data on binary and ternary alloys. Particular attention has been devoted to Fe$_{56}$Cr$_{21}$Ni$_{23}$, Fe$_{38}$Cr$_{14}$Ni$_{48}$ and Fe$_{34}$Cr$_{20}$Ni$_{46}$ alloys close to the centre of the composition triangle, to rationalize how SRO varies in Fe-Cr, Fe-Ni and Ni-Cr binary alloys at various temperatures. The fact that SRO decreases significantly for Fe-Ni pairs as a function of Cr concentration agrees with experimental observations. This important aspect of alloy thermodynamics is also related to the fact that interaction between Cr and both Fe and Ni is strongly anti-ferromagnetic, explaining large negative values of SRO predicted for Fe-Cr and Ni-Cr atomic pairs. Our study shows that it is now possible to treat thermodynamics of Fe-Cr-Ni starting from first principles, taking into account both chemical and magnetic interactions in this traditionally important but complex ternary magnetic alloy. By comparing MC configurations generated using effective cluster interactions with those created by the SQS method, we are able to demonstrate that the former are energetically more stable for all the magnetic structures considered here, as illustrated by the case of Fe$_{70}$Cr$_{20}$Ni$_{10}$ alloy. This provides vital information about the reference structures required for modelling point defects in ternary alloys, where defect properties exhibit high sensitivity not only to the average alloy composition but also to the local chemical and magnetic environment of a defect site \cite{Nguyen-Manh2012,Muzyk2011,Muzyk2013}. \begin{acknowledgments} This work was funded by the Accelerated Metallurgy Project, which is co-funded by the European Commission in the 7th Framework Programme (Contract NMP4-LA-2011-263206), by the European Space Agency and by the individual partner organizations. This work was also part-funded by the RCUK Energy Programme (Grant Number EP/I501045) and by the European Union's Horizon 2020 research and innovation programme under grant agreement number 633053. To obtain further information on the data and models underlying this paper please contact <EMAIL>. The views and opinions expressed herein do not necessarily reflect those of the European Commission. The authors would like to thank Charlotte Becquart, Maria Ganchenkova and George Smith for stimulating and helpful discussions. DNM would like to acknowledge the Juelich supercomputer centre for the provision of High-Performances Computer for Fusion (HPC-FF) facilities as well as the International Fusion Energy Research Centre (IFERC) for the provision of a supercomputer (Helios) at the Computational Simulation Centre (CSC) in Rokkasho (Japan). \end{acknowledgments}
\section{DFS for the sub-Ohmic baths} In the zeroth order approximation, we only select the first term Eq. (\ref {sub-Ohmic}). Similar to the derivation in the symmetric case, projecting the Schr\"{o}dinger equation in the upper level onto the orthogonal basis \left\langle 0\right| \ D^{\dagger }\left( \alpha _k\right) \;$and$\ \left\langle 0\right| a_kD^{\dagger }\left( \alpha _k\right) $ and low level onto $\left\langle 0\right| \ D^{\dagger }\left( \beta _k\right) \;$and$\ \left\langle 0\right| a_kD^{\dagger }\left( \beta _k\right) $ result in \begin{eqnarray} \sum_k\left( \omega _k\alpha _k^2+g_k\alpha _k\right) -\frac \Delta 2r\ \Gamma &=&E, \label{E_upper} \\ \omega _k\alpha _k+\frac 12g_k+\frac \Delta 2r\Gamma D_k &=&0, \label{dis_upper} \end{eqnarray} and \begin{eqnarray} \sum_k\left( \omega _k\beta _k^2-g_k\beta _k\right) -\frac \Delta {2r}\ \Gamma &=&E, \label{E_down} \\ \omega _k\beta _k-\frac 12g_k-\frac \Delta {2r}\Gamma D_k &=&0, \label{dis_down} \end{eqnarray} where \begin{eqnarray*} \Gamma &=&\exp \left[ -\frac 12\sum_kD_k^2\right] , \\ D_k &=&\alpha _k-\beta _k, \end{eqnarray*} which are the same as those obtained variationally within the GSH ansatz\ \cite{Chin}. For the second-order DFS, the first two terms in Eq. (\ref{sub-Ohmic}) is kept. Proceeding as procedures outlines above, Projecting the Schr\"{o}dinger equation in the upper level onto the orthogonal states \left\langle 0\right| \ D^{\dagger }\left( \alpha _k\right) ,\ \left\langle 0\right| a_kD^{\dagger }\left( \alpha _k\right) $, and $\ \left\langle 0\right| a_{k_1}a_{k_2}D^{\dagger }\left( \alpha _k\right) $ and low level onto $\left\langle 0\right| \ D^{\dagger }\left( \beta _k\right) ,\left\langle 0\right| a_kD^{\dagger }\left( \beta _k\right)$, and $\ \left\langle 0\right| a_{k_1}a_{k_2}D^{\dagger }\left( \beta _k\right) $ yield the following six equations \begin{widetext} \begin{eqnarray} \sum_k\left[ \omega _k\alpha _k^2+g_k\alpha _k\right] -\frac \Delta 2\Gamma \left[ r+\sum_kB_kD_k\right] \ &=&E, \label{sub_1up} \\ r\sum_k\left[ \omega _k\beta _k^2-g_k\beta _k\right] -\frac \Delta 2\Gamma \left[ 1+\sum_kA_kD_k\right] \ &=&rE, \label{sub_1down} \end{eqnarray} \begin{equation} \left[ \omega _k\alpha _k+\frac{g_k}2\right] +\sum_{k^{\prime }}2b_1\left( k,k^{\prime }\right) \left[ \omega _{k^{\prime }}\alpha _{k^{\prime }}+\frac g_{k^{\prime }}}2\right] -\Delta \Gamma B_k+\frac \Delta 2\Gamma D_k\left[ r+\sum_kB_kD_k\right] =0, \label{sub_2up} \end{equation} \begin{equation} r\left[ \omega _k\beta _k-\frac{g_k}2\right] +\sum_{k^{\prime }}2b_2\left( k,k^{\prime }\right) \left[ \omega _{k^{\prime }}\beta _{k^{\prime }}-\frac g_{k^{\prime }}}2\right] +\Delta \Gamma A_k-\frac \Delta 2\Gamma D_k\left[ 1+\sum_kA_kD_k\right] =0, \label{sub_2down} \end{equation} \begin{eqnarray} &&b_1(k_1,k_2)\left( \omega _{k_1}+\omega _{k_2}\right) +\frac \Delta 2\Gamma \left[ r+\sum_kB_kD_k\right] b_1(k_1,k_2) \nonumber \\ &&-\frac \Delta 2b_2(k_1,k_2)\Gamma +\frac \Delta 2\Gamma \left[ B_{k_1}D_{k_2}+B_{k_2}D_{k_1}\right] -\frac \Delta 4\Gamma D_{k_1}D_{k_2}\left[ r+\sum_kB_kD_k\right] =0, \label{sub_3up} \end{eqnarray} \begin{eqnarray} &&b_2(k_1,k_2)\left( \omega _{k_1}+\omega _{k_2}\right) +\frac \Delta {2r}\Gamma \left[ 1+\sum_kA_kD_k\right] b_2(k_1,k_2) \nonumber \\ &&-\frac \Delta 2b_1(k_1,k_2)\Gamma +\frac \Delta 2\Gamma \left[ A_{k_1}D_{k_2}+A_{k_2}D_{k_1}\right] -\frac \Delta 4\Gamma D_{k_1}D_{k_2}\left[ 1+\sum_kA_kD_k\right] =0, \label{sub_3down} \end{eqnarray} \end{widetext} where \begin{eqnarray*} \ \ A_k &=&\sum_{k^{\prime }}b_1(k^{\prime },k)D_{k^{\prime }}, \\ \ \ B_k &=&\sum_{k^{\prime }}b_2(k^{\prime },k)D_{k^{\prime }}. \end{eqnarray*} The self-consistent solutions for the four coupled equations Eqs. (\ref {sub_2up}), (\ref{sub_2down}), (\ref{sub_3up}) and (\ref{sub_3down}) will give all results in the second-order DFS. If set $r=1,\alpha _k=-\beta _k$ and $b_1(k_1,k_2)=b_2(k_1,k_2)$, Eqs. (\ref{Eq_21}) and (\ref{Eq_22}) in the symmetric case are recovered completely. \section{Gaussian-logarithmical integration for the continuous integral} The symmetrical case is also used to illustrate a effective numerical approach to the calculation of the summation clearly. In the zeroth-order approximation, also the well known SH ansatz, we can set \[ \alpha _k=\alpha _k^{\prime }g_k, \] Eq. (\ref{Eq_dis}) becomes \[ \alpha _k^{\prime }=-\frac{1/2}{\omega _k+\Delta \exp \left( -2\sum_k\alpha _k^{\prime 2}g_k^2\right) }, \] so $\alpha _k^{\prime }$ is only related to $g_k$ implicitly. According to the spectral density, we have \begin{equation} \alpha ^{\prime }(\omega )=-\frac{1/2}{\omega +\Delta \exp \left[ -\frac 2\pi \int_0^{\omega _c}d\omega ^{\prime }\alpha ^{\prime 2}(\omega ^{\prime })J(\omega ^{\prime })\right] }, \label{SH_cont} \end{equation} which can be solved numerically by iterations. In the second-order approximation, we can set \begin{eqnarray*} \alpha _k &=&\alpha _k^{\prime }g_k, \\ b_{k_1,k_2} &=&b_{k_1,k_2}^{\prime }g_{_{k_1}}g_{_{k_2}}. \end{eqnarray*} Inserting to Eqs. (\ref{Eq_21}) and (\ref{Eq_22}) gives \[ \alpha _k^{\prime }=\frac{-\frac 12+2\sum_{k^{\prime }}g_{k^{\prime }}^2b_{k,k^{\prime }}^{\prime }\left[ \left( \omega _{k^{\prime }}-\Delta \eta \right) \alpha _{k^{\prime }}^{\prime }+1/2\right] }{\omega _k+\Delta \eta \left( 1+4\zeta \right) }, \] \[ b_{k_1,k_2}^{\prime }=\frac{\alpha _{k_1}^{\prime }\alpha _{k_2}^{\prime }\left( 1+4\zeta \right) -\sum_{k^{\prime }}g_{k^{\prime }}^2\alpha _{k^{\prime }}^{\prime }\left( b_{k_1,k^{\prime }}^{\prime }\alpha _{k_2}^{\prime }+b_{k_2,k^{\prime }}^{\prime }\alpha _{k_1}^{\prime }\right) }{2\zeta +\left( \omega _{k_1}+\omega _{k_2}\right) /\left( \Delta \eta \right) }, \] where \[ \zeta =\sum_kg_k^2\sum_{k^{\prime }}g_{k^{\prime }}^2b_{k,k^{\prime }}^{\prime }\alpha _{k^{\prime }}^{\prime }\alpha _k^{\prime }. \] Given $g_k$, both $\alpha _k^{\prime }$ and $b_{k_1,k_2}^{\prime }\;$can be obtained self-consistently. Note that each $k$-summation takes the form of \sum_kg_k^2I(k)$ where $I(k)$ does not depend on $g_k^2\;$explicitly, and so both $\alpha _k^{\prime }$ and $b_{k_1,k_2}^{\prime }\;$are functionals of g_k$. $\;$Without loss of generality, $k$ is corresponding to $\omega $ one by one, the $k$-summation can be transformed to the $\omega \;$integral as \[ \sum_kg_k^2I(k)\rightarrow \int_0^{\omega _c}d\omega \frac{J(\omega )}\pi I(\omega ), \] so we have \begin{equation} \alpha ^{\prime }(\omega )=\frac{-\frac 12+\xi (\omega )-2\Delta \eta \chi (\omega )}{\omega +\Delta \eta \left( 1+4\zeta \right) }, \label{displacement} \end{equation} \begin{equation} b^{\prime }\left( \omega _1,\omega _2\right) =\frac{\alpha ^{\prime }(\omega _1)\alpha ^{\prime }(\omega _2)\left( 1+4\zeta \right) -\kappa (\omega _1,\omega _2)}{2\zeta +\left( \omega _{_1}+\omega _{_2}\right) /(\Delta \eta )}, \label{sec_coeff} \end{equation} where \begin{eqnarray*} \xi (\omega ) &=&\int_0^{\omega _c}d\omega ^{\prime }\frac{J(\omega ^{\prime })}\pi \left[ 2\omega ^{\prime }\alpha ^{\prime }(\omega ^{\prime })+1\right] b^{\prime }\left( \omega ,\omega ^{\prime }\right) , \\ \chi (\omega ) &=&\int_0^{\omega _c}d\omega ^{\prime }\frac{J(\omega ^{\prime })}\pi \alpha ^{\prime }(\omega ^{\prime })b^{\prime }\left( \omega ,\omega ^{\prime }\right) , \\ \kappa (\omega _1,\omega _2) &=&\chi (\omega _1)\alpha ^{\prime }(\omega _2)+\chi (\omega _2)\alpha ^{\prime }(\omega _1), \end{eqnarray*} are some functions for $\omega ,$ and \begin{eqnarray*} \zeta &=&\int_0^{\omega _c}d\omega \frac{J(\omega )}\pi \int_0^{\omega _c}d\omega ^{\prime }\frac{J(\omega ^{\prime })}\pi \alpha ^{\prime }(\omega )\alpha ^{\prime }(\omega ^{\prime })b^{\prime }\left( \omega ,\omega ^{\prime }\right) , \\ \eta &=&\exp \left[ -2\int_0^{\omega _c}d\omega ^{\prime }\frac{J(\omega ^{\prime })}\pi \alpha ^{\prime 2}(\omega ^{\prime })\right] , \end{eqnarray*} are constants. If both $\alpha ^{\prime }(\omega )$ and $b^{\prime }\left( \omega _1,\omega _2\right) $ are obtained, all observables can in turn be calculated. For example, using Eq. (\ref{Eq_2_en}), the energy in the second-order DFS can be calculate as \begin{equation} E=\int_0^{\omega _c}d\omega \frac{J(\omega )}\pi \alpha ^{\prime }(\omega )\left[ \omega \alpha ^{\prime }(\omega )+1\right] -\frac 12\Delta \eta \left( 1+4\zeta \right) \ . \end{equation} The self-consistent solutions in the coupled equations Eqs. (\ref {displacement}) and (\ref{sec_coeff})) are in no way obtained analytically, numerical calculation should be performed. Note that the low frequency modes play the dominant role in the QPT of the sub-ohmic spin-boson model. At the critical point, there is an infrared divergence of the integrand like \int_0^{\omega _c}\omega ^{s-2}d\omega \;$in the limit of $\omega \rightarrow 0$ for sub-ohmic bath, which is called as the infrared catastrophe. Thanks to the Gaussian quadrature rules, where the zero frequency is not touched. We can discretize the whole frequency interval with Gaussian grids, the integral can be numerically exactly achieved with a large number of Gaussian grids. It is very time consuming to calculate the integral in this way, especially for high dimensional integral involved in the high-order DFS. According to the structure of the integrand, it is not economical to deal with the high and low frequency regime on the equal footing. To increase the efficiency, we combine the logarithmic discretization and Gaussian quadrature rule. First, we divide the $\omega $ interval $[0,1]$ into $M+1$ sub-intervals as $[\Lambda ^{-(m+1)},\Lambda ^{-m}]\;(m=0,1,2,M-1)\;$and$\;[0,\Lambda ^{-M}]$ , then we apply the Gaussian quadrature rule to each logarithmical sub-interval. So the continuous integral is calculated by the following summation \begin{equation} \int_0^1J(\omega )I(\omega )d\omega =\sum_{m=0}^M\sum_{n=1}^NW_{m,n}J(\omega _{m,n})I(\omega _{m,n}), \label{combination} \end{equation} where $N$ is the number of gaussian points inserted in each sub-interval, W_{m,n}$ is corresponding Gaussian weight. \begin{figure}[tbp] \includegraphics[width=8cm]{Compare_GSH.eps} \caption{ Magnetization $\langle \sigma _z\rangle$ as a function of the coupling strength $\lambda$ in the GSH ansatz. (a) For $s = 0.6$, converged results within GL integration (open squares), numerical exact ones (solid lines), and those within logarithmic discretization with different truncation numbers $K = 10, 20$, and $30$. (b) The converged magnetization within GL integration (open circles) and logarithmic discretization(filled squares) for $s = 0.2,0.4$, and $0.6$. Numerical exact ones are denoted by the solid lines.} \label{CompareIntegral} \end{figure} \begin{figure}[tbp] \includegraphics[width=8cm]{converge_test.eps} \caption{ (Color online) Magnetization $\langle \sigma _z\rangle$ as a function of the coupling strength $\lambda$ in the second-order DFS within GL integration using different $M, N$, and $\Lambda$ for (a) $s=0.6$ and (b) $s=0.8$(b).} \label{Fig_converge} \end{figure} To demonstrate the efficiency of the Gaussian-logarithmical (GL) integration, we first apply it to the GSH ansatz, which is also the zero-order approximation in the DFS. The one-dimensional integral can be numerically exactly done by Gaussian integration over the whole interval with a huge number of discretizations, and corresponding results can be regarded as a benchmark. After careful examinations, using the GL technique, the converging results for the magnetization can be archived if set $M=6,N=9, $ and $\Lambda =9$ The corresponding results for $s=0.6$ are presented in Fig. \ref{CompareIntegral} (a) with open squares, which agrees excellently with the numerically exact one by a huge number of discretization in the Gaussian integration (solid lines). We can also perform the logarithmic discretization of the bosonic energy band, as was widely used in the previous studies, such as NRG \cite{Bulla} and multi-coherent states \cite{mD1}. In the GSH ansatz, this can be easily done by set \begin{equation} g_k^2=\int_{\Lambda ^{-(k+1)}}^{\Lambda ^{-k}}\frac{J(\omega )}\pi d\omega ,\;\omega _k=\frac 1{g_k^2}\int_{\Lambda ^{-(k+1)}}^{\Lambda ^{-k}}\frac J(\omega )}\pi \omega d\omega , \label{Dis_GSH_H} \end{equation} in Eqs. (A1-A4) of Appendix A. The spectral density is truncated to a number $K$ of modes. The summation is performed over the integer $k$ directly and the self-consistent solution with discretized form can be also obtained. The logarithmic grid is chosen as $\Lambda =2$, the same as that in Refs. \cite {Bulla,mD1}. The magnetization as a function of $\lambda $ for $s=0.6$ with such a logarithmic discretization are collected in Fig. \ref{CompareIntegral} (a) with different truncation number $K$ of bosonic modes. The converging results can be also obtained for $K\geq 20$, which is however obviously different from the numerically exact one. Note that this kind of logarithmic discretization of the bosonic energy band at the very beginning is not equivalent to the logarithmic discretization of the continuous integral derived in the end of the DFS approach. The converged magnetization within both the GL and logarithmic discreatization for different values of $s$ are carefully examined, and the results are exhibited in Fig. \ref{CompareIntegral} (b). The deviation between these two convergent ones increases with $s$, and becomes remarkable for $s\ge 0.3$. Then we turn to the second-order DFS study, a central issue in this work. Two-dimensional integral will be involved in this case, so direct Gaussian integral with huge number of discretization is practically difficult. Fortunately, it has been convincingly shown above that in the framework of GSH ansatz, the Gaussian-logarithmic discretization with dozens of grids to the continuous integral can effectively give results with very high accuracy. Therefore we extend this numerical technique to the present case. Interestingly, a excellent convergence behavior for $s=0.6$(a) and $s=0.8 (b) is demonstrated in Fig. \ref{Fig_converge} with different value of $M,N , and $\Lambda $. The converged results obtained in this way compose the main achievement in this work.
\section{Introduction} \label{introSec} Subspace identification arises in a wide variety of signal and information processing applications. In many cases, especially high-dimensional situations, it is common to encounter missing data. Hence the growing literature concerning the estimation of low-dimensional subspaces and matrices from incomplete data in theory \cite{balzano, chi, mardani,candes,recht,kiraly,jain} and applications \cite{eriksson,he}. This paper considers the problem of identifying an \phantomsection\label{rDef}${\hyperref[rDef]{r}}$-dimensional subspace of \phantomsection\label{dDef}$\mathbb{R}^{\fix{{\hyperref[dDef]{d}}}}$ from projections of the subspace onto small subsets of the canonical coordinates. The main contribution of this paper is to establish deterministic necessary and sufficient conditions on such subsets that guarantee that there is {\em only} one ${\hyperref[rDef]{r}}$-dimensional subspace consistent with all the projections. These conditions also have implications for low-rank matrix completion and related problems. \subsection*{Organization of the paper} In {\hyperref[modelSec]{Section \ref{modelSec}}}\ we formally state the problem and our main results. We present the proof of our main theorem in {\hyperref[proofSec]{Section \ref{proofSec}}}. {\hyperref[LRMCSec]{Section \ref{LRMCSec}}}\ illustrates the implications of our results for low-rank matrix completion. {\hyperref[graphSec]{Section \ref{graphSec}}}\ presents the graphical interpretation of the problem and another necessary condition based on this viewpoint. \vspace{.1cm} \section{Model and main results} \label{modelSec} \vspace{.1cm} Let \phantomsection\label{sstarDef}${\hyperref[sstarDef]{S^\star}}$ denote an ${\hyperref[rDef]{r}}$-dimensional subpace of $\mathbb{R}^{\fix{{\hyperref[dDef]{d}}}}$. Define \phantomsection\label{OODef}${\hyperref[OODef]{\bs{\Omega}}}$ as a \phantomsection\label{NDef}${\hyperref[dDef]{d}} \times {\hyperref[NDef]{N}}$ binary matrix and let \phantomsection\label{oiDef}${\hyperref[oiDef]{\bs{\omega}_i}}$ denote the \phantomsection\label{iDef}${\hyperref[iDef]{i}}^{th}$ column of ${\hyperref[OODef]{\bs{\Omega}}}$. The nonzero entries of ${\hyperref[oiDef]{\bs{\omega}_i}}$ indicate the canonical coordinates involved in the ${\hyperref[iDef]{i}}^{th}$ projection. Since ${\hyperref[sstarDef]{S^\star}}$ is ${\hyperref[rDef]{r}}$-dimensional, the restriction of ${\hyperref[sstarDef]{S^\star}}$ onto ${\hyperref[LDef]{\ell}} \leq {\hyperref[rDef]{r}}$ coordinates will be $\mathbb{R}^{\fix{{\hyperref[LDef]{\ell}}}}$ (in general), and hence such a projection will provide no information specific to ${\hyperref[sstarDef]{S^\star}}$. Therefore, without loss of generality (see the {\hyperref[generalizationApx]{appendix}}\ for immediate generalizations) we will assume that: \begin{itemize} \phantomsection\label{AoneAssDef} \item[{\hyperref[AoneAssDef]{\textbf{A1}}}] ${\hyperref[OODef]{\bs{\Omega}}}$ has exactly ${\hyperref[rDef]{r}}+1$ nonzero entries per column. \end{itemize} Given an ${\hyperref[rDef]{r}}$-dimensional subspace \phantomsection\label{sDef}${\hyperref[sDef]{S}}$, let \phantomsection\label{soiDef}${\hyperref[soiDef]{S_{\bs{\omega}_i}}} \subset \mathbb{R}^{\fix{{\hyperref[rDef]{r}}}+1}$ denote the restriction of ${\hyperref[sDef]{S}}$ to the nonzero coordinates in ${\hyperref[oiDef]{\bs{\omega}_i}}$. The question addressed in this paper is whether the restrictions $\{{\hyperref[soiDef]{S^\star_{\bs{\omega}_i}}}\}_{\fix{{\hyperref[iDef]{i}}}=1}^{\fix{{\hyperref[NDef]{N}}}}$ uniquely determine ${\hyperref[sstarDef]{S^\star}}$. This depends on the sampling pattern in ${\hyperref[OODef]{\bs{\Omega}}}$. \begin{figure} \centering \includegraphics[width=3.45cm]{projection} \label{illustration} \caption{When can ${\hyperref[sstarDef]{S^\star}}$ be identified from its canonical projections $\{{\hyperref[soiDef]{S^\star_{\bs{\omega}_i}}}\}_{\fix{{\hyperref[iDef]{i}}}=1}^{\fix{{\hyperref[NDef]{N}}}}$?} \end{figure} We will see that identifiability of this sort can only be possible if ${\hyperref[NDef]{N}} \geq {\hyperref[dDef]{d}}-{\hyperref[rDef]{r}}$, since $\ker {\hyperref[sstarDef]{S^\star}}$ is $({\hyperref[dDef]{d}}-{\hyperref[rDef]{r}})$-dimensional. Thus, unless otherwise stated, we will also assume that: \begin{itemize} \phantomsection\label{AtwoAssDef} \item[{\hyperref[AtwoAssDef]{\textbf{A2}}}] ${\hyperref[OODef]{\bs{\Omega}}}$ has exactly ${\hyperref[NDef]{N}}={\hyperref[dDef]{d}}-{\hyperref[rDef]{r}}$ columns. \end{itemize} Let \phantomsection\label{GrDef}$\hyperref[GrDef]{{\rm Gr}}({\hyperref[rDef]{r}},\mathbb{R}^{\fix{{\hyperref[dDef]{d}}}})$ denote the Grassmannian manifold of ${\hyperref[rDef]{r}}$-dimensional subspaces in $\mathbb{R}^{\fix{{\hyperref[dDef]{d}}}}$. Define \phantomsection\label{SSDef}${\hyperref[SSDef]{\mathscr{S}}}({\hyperref[sstarDef]{S^\star}},{\hyperref[OODef]{\bs{\Omega}}})$ $\subset$ $\hyperref[GrDef]{{\rm Gr}}({\hyperref[rDef]{r}},\mathbb{R}^{\fix{{\hyperref[dDef]{d}}}})$ such that every ${\hyperref[sDef]{S}} \in {\hyperref[SSDef]{\mathscr{S}}}({\hyperref[sstarDef]{S^\star}},{\hyperref[OODef]{\bs{\Omega}}})$ satisfies ${\hyperref[soiDef]{S_{\bs{\omega}_i}}}={\hyperref[soiDef]{S^\star_{\bs{\omega}_i}}}$ $\forall$ ${\hyperref[iDef]{i}}$. In words, ${\hyperref[SSDef]{\mathscr{S}}}({\hyperref[sstarDef]{S^\star}},{\hyperref[OODef]{\bs{\Omega}}})$ is the set of all ${\hyperref[rDef]{r}}$-dimensional subspaces matching ${\hyperref[sstarDef]{S^\star}}$ on ${\hyperref[OODef]{\bs{\Omega}}}$. \begin{myExample} \label{introEg} Let ${\hyperref[dDef]{d}}=5$, ${\hyperref[rDef]{r}}=1$, \begin{align*} {\hyperref[sstarDef]{S^\star}} \ = \ {\rm span}\left[\begin{matrix} 1 \\ 2 \\ 3 \\ 4 \\ 4 \end{matrix}\right] \hspace{.25cm} \text{and } \hspace{.25cm} {\hyperref[OODef]{\bs{\Omega}}} \ = \ \left[\begin{matrix} 1 & 0 & 1 & 0 \\ 1 & 1 & 0 & 0 \\ 0 & 1 & 1 & 0 \\ 0 & 0 & 0 & 1 \\ 0 & 0 & 0 & 1 \\ \end{matrix}\right]. \end{align*} Then, for example, \begin{align*} \hyperref[soiDef]{S^\star_{\bs{\omega}_3}}={\rm span}\left[\begin{matrix} 1 \\ 3\end{matrix}\right]. \end{align*} It is easy to see that there are infinitely many $1$-dimensional subspaces that match ${\hyperref[sstarDef]{S^\star}}$ on ${\hyperref[OODef]{\bs{\Omega}}}$. In fact, \begin{align*} {\hyperref[SSDef]{\mathscr{S}}}({\hyperref[sstarDef]{S^\star}},{\hyperref[OODef]{\bs{\Omega}}}) \ = \ \bigg\{ {\rm span}[1 \ \ 2 \ \ 3 \ \ \alpha \ \ \alpha]^\mathsf{T} \ : \ \alpha \in \mathbb{R} \backslash \{0\} \bigg\}. \end{align*} However, if we instead had $\hyperref[oiDef]{\bs{\omega}_3}=[0 \ \ 0 \ \ 1 \ \ 1 \ \ 0]^\mathsf{T}$, then ${\hyperref[sstarDef]{S^\star}}$ would be the only subspace in ${\hyperref[SSDef]{\mathscr{S}}}({\hyperref[sstarDef]{S^\star}},{\hyperref[OODef]{\bs{\Omega}}})$. \end{myExample} The main result of this paper is the following theorem, which gives necessary and sufficient conditions on ${\hyperref[OODef]{\bs{\Omega}}}$ to guarantee that ${\hyperref[SSDef]{\mathscr{S}}}({\hyperref[sstarDef]{S^\star}},{\hyperref[OODef]{\bs{\Omega}}})$ contains no subspace other than ${\hyperref[sstarDef]{S^\star}}$. Our results hold for \phantomsection\label{aeDef}{\rm({\hyperref[aeDef]{{\rm a.e.}}})} ${\hyperref[sstarDef]{S^\star}}$, with respect to the uniform measure over $\hyperref[GrDef]{{\rm Gr}}({\hyperref[rDef]{r}},\mathbb{R}^{\fix{{\hyperref[dDef]{d}}}})$. Given a matrix, let $\phantomsection\label{nOfDef}{\hyperref[nOfDef]{n}}(\bs{\cdot})$ denote its number of columns, and \phantomsection\label{mOfDef}${\hyperref[mOfDef]{m}}(\bs{\cdot})$ the number of its {\em nonzero} rows. \begin{framed} \begin{myTheorem} \label{identifiabilityThm} Let {\hyperref[AoneAssDef]{\textbf{A1}}}\ and {\hyperref[AtwoAssDef]{\textbf{A2}}}\ hold. For \hyperref[aeDef]{almost every} ${\hyperref[sstarDef]{S^\star}}$, ${\hyperref[sstarDef]{S^\star}}$ is the only subspace in ${\hyperref[SSDef]{\mathscr{S}}}({\hyperref[sstarDef]{S^\star}},{\hyperref[OODef]{\bs{\Omega}}})$ if and only if every matrix \phantomsection\label{ODef}${\hyperref[ODef]{\bs{\Omega}'}}$ formed with a subset of the columns in ${\hyperref[OODef]{\bs{\Omega}}}$ satisfies \begin{align} \label{identifiabilityEq} {\hyperref[mOfDef]{m}}({\hyperref[ODef]{\bs{\Omega}'}}) \ \geq \ {\hyperref[nOfDef]{n}}({\hyperref[ODef]{\bs{\Omega}'}}) + {\hyperref[rDef]{r}}. \end{align} \end{myTheorem} \end{framed} The proof of {\hyperref[identifiabilityThm]{Theorem \ref{identifiabilityThm}}}\ is given in {\hyperref[proofSec]{Section \ref{proofSec}}}. In words, {\hyperref[identifiabilityThm]{Theorem \ref{identifiabilityThm}}}\ is stating that ${\hyperref[sstarDef]{S^\star}}$ is the only subspace that matches ${\hyperref[sstarDef]{S^\star}}$ in ${\hyperref[OODef]{\bs{\Omega}}}$ if and only if every subset of ${\hyperref[nOfDef]{n}}$ columns of ${\hyperref[OODef]{\bs{\Omega}}}$ has at least ${\hyperref[nOfDef]{n}}+{\hyperref[rDef]{r}}$ nonzero rows. \begin{myExample} The following matrix, where \phantomsection\label{oneDef}${\hyperref[oneDef]{\bs{{\rm 1}}}}$ denotes a block of all $1$'s and \phantomsection\label{IDef}${\hyperref[IDef]{\bs{{\rm I}}}}$ denotes the identity matrix, satisfies the conditions of {\hyperref[identifiabilityThm]{Theorem \ref{identifiabilityThm}}}: \begin{align*} {\hyperref[OODef]{\bs{\Omega}}} \ = \ \left[ \begin{array}{c} \hspace{.3cm} \Scale[1.5]{{\hyperref[oneDef]{\bs{{\rm 1}}}}} \hspace{.3cm} \\ \hline \\ \Scale[1.5]{{\hyperref[IDef]{\bs{{\rm I}}}}} \\ \\ \end{array}\right] \begin{matrix} \left. \begin{matrix} \\ \end{matrix} \right\} {\hyperref[rDef]{r}} \hspace{.7cm} \\ \left. \begin{matrix} \\ \\ \\ \end{matrix} \right\} {\hyperref[dDef]{d}}-{\hyperref[rDef]{r}}. \end{matrix} \end{align*} \end{myExample} When the conditions of {\hyperref[identifiabilityThm]{Theorem \ref{identifiabilityThm}}}\ are satisfied, identifying ${\hyperref[sstarDef]{S^\star}}$ becomes a trivial task: ${\hyperref[sstarDef]{S^\star}}= \ker {\hyperref[AADef]{\bs{{\rm A}}}}{}^\mathsf{T}$, with ${\hyperref[AADef]{\bs{{\rm A}}}}$ as defined in {\hyperref[proofSec]{Section \ref{proofSec}}}. In general, verifying the conditions on ${\hyperref[OODef]{\bs{\Omega}}}$ in {\hyperref[identifiabilityThm]{Theorem \ref{identifiabilityThm}}}\ may be computationally prohibitive, especially for large ${\hyperref[dDef]{d}}$. However, as the next theorem states, uniform random sampling patterns will satisfy the conditions in {\hyperref[identifiabilityThm]{Theorem \ref{identifiabilityThm}}}\ with high probability \phantomsection\label{whpDef}({\hyperref[whpDef]{w.h.p}}.). \begin{myTheorem} \label{probabilityThm} Assume {\hyperref[AtwoAssDef]{\textbf{A2}}}\ and let \phantomsection\label{epsDef}$0<{\hyperref[epsDef]{\epsilon}} \leq 1$ be given. Suppose ${\hyperref[rDef]{r}} \leq \frac{\fix{{\hyperref[dDef]{d}}}{}}{6}$ and that each column of ${\hyperref[OODef]{\bs{\Omega}}}$ contains at least \phantomsection\label{LDef}${\hyperref[LDef]{\ell}}$ nonzero entries, selected uniformly at random and independently across columns, with \begin{align} \label{kEq} \textstyle {\hyperref[LDef]{\ell}} \ \geq \ \max \left\{9 \log(\frac{\fix{{\hyperref[dDef]{d}}}{}}{\fix{{\hyperref[epsDef]{\epsilon}}}{}})+12, \ 2{\hyperref[rDef]{r}} \right\}. \end{align} Then ${\hyperref[OODef]{\bs{\Omega}}}$ will satisfy the conditions of {\hyperref[identifiabilityThm]{Theorem \ref{identifiabilityThm}}}\ with probability at least $1-{\hyperref[epsDef]{\epsilon}}$. \end{myTheorem} {\hyperref[probabilityThm]{Theorem \ref{probabilityThm}}}\ is proved in the {\hyperref[probabilityApx]{appendix}}. Notice that $\mathscr{O}({\hyperref[rDef]{r}} \log{\hyperref[dDef]{d}})$ nonzero entries per column is a typical requirement of {\hyperref[LRMCDef]{LRMC}}\ methods, while $\mathscr{O}(\max\{{\hyperref[rDef]{r}},\log{\hyperref[dDef]{d}}\})$ is sufficient for subspace identifiability. \vspace{.3cm} \section{Proof of {\hyperref[identifiabilityThm]{Theorem \ref{identifiabilityThm}}}} \label{proofSec} \vspace{.3cm} For any subspace, matrix or vector that is compatible with a binary vector \phantomsection\label{oDef}${\hyperref[oDef]{\bs{\upsilon}}}$, we will use the subscript \phantomsection\label{cdotoDef}${\hyperref[oDef]{\bs{\upsilon}}}$ to denote its restriction to the nonzero coordinates/rows in ${\hyperref[oDef]{\bs{\upsilon}}}$. For {\hyperref[aeDef]{{\rm a.e.}}}\ ${\hyperref[sstarDef]{S^\star}}$, ${\hyperref[soiDef]{S^\star_{\bs{\omega}_i}}}$ is an ${\hyperref[rDef]{r}}$-dimensional subspace of $\mathbb{R}^{\fix{{\hyperref[rDef]{r}}}+1}$, and the kernel of ${\hyperref[soiDef]{S^\star_{\bs{\omega}_i}}}$ is a $1$-dimensional subspace of $\mathbb{R}^{\fix{{\hyperref[rDef]{r}}}+1}$. \begin{myLemma} \label{aEntriesLem} Let \phantomsection\label{aoiDef}${\hyperref[aoiDef]{\bs{{\rm a}}_{\bs{\omega}_i}}} \in \mathbb{R}^{\fix{{\hyperref[rDef]{r}}}+1}$ be a nonzero element of $\ker {\hyperref[soiDef]{S^\star_{\bs{\omega}_i}}}$. All entries of ${\hyperref[aoiDef]{\bs{{\rm a}}_{\bs{\omega}_i}}}$ are nonzero for {\hyperref[aeDef]{{\rm a.e.}}}\ ${\hyperref[sstarDef]{S^\star}}$. \end{myLemma} \begin{proof} Suppose ${\hyperref[aoiDef]{\bs{{\rm a}}_{\bs{\omega}_i}}}$ has at least one zero entry. Use ${\hyperref[oDef]{\bs{\upsilon}}}$ to denote the binary vector of the nonzero entries of ${\hyperref[aoiDef]{\bs{{\rm a}}_{\bs{\omega}_i}}}$. Since ${\hyperref[aoiDef]{\bs{{\rm a}}_{\bs{\omega}_i}}}$ is orthogonal to ${\hyperref[soiDef]{S^\star_{\bs{\omega}_i}}}$, for every $\hyperref[cdotoDef]{\bs{{\rm u}}_{\bs{\omega}_i}} \in {\hyperref[soiDef]{S^\star_{\bs{\omega}_i}}}$ we have that ${\hyperref[aoiDef]{\bs{{\rm a}}^\T_{\bs{\omega}_i}}} \hyperref[cdotoDef]{\bs{{\rm u}}_{\bs{\omega}_i}}={\hyperref[cdotoDef]{\bs{{\rm a}}^\T_{\bs{\upsilon}}}} \hyperref[cdotoDef]{\bs{{\rm u}}_{\bs{\upsilon}}}=0$. Then \phantomsection\label{sstaroDef}${\hyperref[cdotoDef]{S^\star_{\bs{\upsilon}}}}$ satisfies \begin{align} \label{aEntriesEq} \dim {\hyperref[cdotoDef]{S^\star_{\bs{\upsilon}}}} \ \leq \ \dim \ker {\hyperref[cdotoDef]{\bs{{\rm a}}^\T_{\bs{\upsilon}}}} \ = \ \|{\hyperref[oDef]{\bs{\upsilon}}}\|_1-1 \ < \ \|{\hyperref[oDef]{\bs{\upsilon}}}\|_1. \end{align} Observe that for every binary vector ${\hyperref[oDef]{\bs{\upsilon}}}$ with $\|{\hyperref[oDef]{\bs{\upsilon}}}\|_1 \leq {\hyperref[rDef]{r}}$, {\hyperref[aeDef]{{\rm a.e.}}}\ ${\hyperref[rDef]{r}}$-dimensional subspace ${\hyperref[sDef]{S}}$ satisfies $\dim {\hyperref[cdotoDef]{S_{\bs{\upsilon}}}} = \|{\hyperref[oDef]{\bs{\upsilon}}}\|_1$. Thus \eqref{aEntriesEq} holds only in a set of measure zero. \end{proof} Define \phantomsection\label{aiDef}${\hyperref[aiDef]{\bs{{\rm a}}_i}}$ as the vector in $\mathbb{R}^{\fix{{\hyperref[dDef]{d}}}}$ with the entries of ${\hyperref[aoiDef]{\bs{{\rm a}}_{\bs{\omega}_i}}}$ in the nonzero positions of ${\hyperref[oiDef]{\bs{\omega}_i}}$ and zeros elsewhere. Then ${\hyperref[sDef]{S}} \subset \ker {\hyperref[aiDef]{\bs{{\rm a}}^\T_i}}$ for every ${\hyperref[sDef]{S}} \in {\hyperref[SSDef]{\mathscr{S}}}({\hyperref[sstarDef]{S^\star}},{\hyperref[OODef]{\bs{\Omega}}})$ and every ${\hyperref[iDef]{i}}$. Letting \phantomsection\label{AADef}${\hyperref[AADef]{\bs{{\rm A}}}}$ be the ${\hyperref[dDef]{d}} \times ({\hyperref[dDef]{d}}-{\hyperref[rDef]{r}})$ matrix formed with $\{{\hyperref[aiDef]{\bs{{\rm a}}_i}}\}_{\fix{{\hyperref[iDef]{i}}}=1}^{\fix{{\hyperref[dDef]{d}}}-\fix{{\hyperref[rDef]{r}}}}$ as columns, we have that ${\hyperref[sDef]{S}} \subset \ker {\hyperref[AADef]{\bs{{\rm A}}}}{}^\mathsf{T}$ for every ${\hyperref[sDef]{S}}\in {\hyperref[SSDef]{\mathscr{S}}}({\hyperref[sstarDef]{S^\star}},{\hyperref[OODef]{\bs{\Omega}}})$. Note that if $\dim \ker {\hyperref[AADef]{\bs{{\rm A}}}}{}^\mathsf{T}={\hyperref[rDef]{r}}$, then ${\hyperref[SSDef]{\mathscr{S}}}({\hyperref[sstarDef]{S^\star}},{\hyperref[OODef]{\bs{\Omega}}})$ contains just one element, ${\hyperref[sstarDef]{S^\star}}$, which is the identifiability condition of interest. Thus, we will establish conditions on ${\hyperref[OODef]{\bs{\Omega}}}$ guaranteeing that the ${\hyperref[dDef]{d}}-{\hyperref[rDef]{r}}$ columns of ${\hyperref[AADef]{\bs{{\rm A}}}}$ are linearly independent. Recall that for any matrix \phantomsection\label{ADef}${\hyperref[ADef]{\bs{{\rm A}}'}}$ formed with a subset of the columns in ${\hyperref[AADef]{\bs{{\rm A}}}}$, ${\hyperref[nOfDef]{n}}({\hyperref[ADef]{\bs{{\rm A}}'}})$ denotes the number of columns in ${\hyperref[ADef]{\bs{{\rm A}}'}}$, and ${\hyperref[mOfDef]{m}}({\hyperref[ADef]{\bs{{\rm A}}'}})$ denotes the number of {\em nonzero} rows in ${\hyperref[ADef]{\bs{{\rm A}}'}}$. \begin{myLemma} \label{independenceLem} For {\hyperref[aeDef]{{\rm a.e.}}}\ ${\hyperref[sstarDef]{S^\star}}$, the columns of ${\hyperref[AADef]{\bs{{\rm A}}}}$ are linearly dependent if and only if ${\hyperref[nOfDef]{n}}({\hyperref[ADef]{\bs{{\rm A}}'}})>{\hyperref[mOfDef]{m}}({\hyperref[ADef]{\bs{{\rm A}}'}})-{\hyperref[rDef]{r}}$ for some matrix ${\hyperref[ADef]{\bs{{\rm A}}'}}$ formed with a subset of the columns in ${\hyperref[AADef]{\bs{{\rm A}}}}$. \end{myLemma} We will show {\hyperref[independenceLem]{Lemma \ref{independenceLem}}}\ using Lemmas \ref{liLem} and \ref{basisLem} below. Let \phantomsection\label{liOfDef}${\hyperref[liOfDef]{\aleph}}({\hyperref[ADef]{\bs{{\rm A}}'}})$ be the largest number of linearly independent columns in ${\hyperref[ADef]{\bs{{\rm A}}'}}$, i.e., the column rank of ${\hyperref[ADef]{\bs{{\rm A}}'}}$. \begin{myLemma} \label{liLem} For {\hyperref[aeDef]{{\rm a.e.}}}\ ${\hyperref[sstarDef]{S^\star}}$, ${\hyperref[liOfDef]{\aleph}}({\hyperref[ADef]{\bs{{\rm A}}'}}) \leq {\hyperref[mOfDef]{m}}({\hyperref[ADef]{\bs{{\rm A}}'}}) - {\hyperref[rDef]{r}}$. \end{myLemma} \begin{proof} Let ${\hyperref[oDef]{\bs{\upsilon}}}$ be the binary vector of nonzero rows of ${\hyperref[ADef]{\bs{{\rm A}}'}}$, and \phantomsection\label{AoDef}${\hyperref[AoDef]{\bs{{\rm A}}'_{\bs{\upsilon}}}}$ be the ${\hyperref[mOfDef]{m}}({\hyperref[ADef]{\bs{{\rm A}}'}}) \times {\hyperref[nOfDef]{n}}({\hyperref[ADef]{\bs{{\rm A}}'}})$ matrix formed with these rows. For {\hyperref[aeDef]{{\rm a.e.}}}\ ${\hyperref[sstarDef]{S^\star}}$, $\dim {\hyperref[cdotoDef]{S^\star_{\bs{\upsilon}}}}={\hyperref[rDef]{r}}$. Since ${\hyperref[cdotoDef]{S^\star_{\bs{\upsilon}}}} \subset \ker {\hyperref[AoDef]{\bs{{\rm A}}'^\T_{\bs{\upsilon}}}}$, ${\hyperref[rDef]{r}}=\dim {\hyperref[cdotoDef]{S^\star_{\bs{\upsilon}}}} \leq \dim \ker {\hyperref[AoDef]{\bs{{\rm A}}'^\T_{\bs{\upsilon}}}} = {\hyperref[mOfDef]{m}}({\hyperref[ADef]{\bs{{\rm A}}'}})-{\hyperref[liOfDef]{\aleph}}({\hyperref[ADef]{\bs{{\rm A}}'}}) $. \end{proof} We say ${\hyperref[ADef]{\bs{{\rm A}}'}}$ is \phantomsection\label{mldDef}{\em {\hyperref[mldDef]{minimally linearly dependent}}} if the columns in ${\hyperref[ADef]{\bs{{\rm A}}'}}$ are linearly dependent, but every proper subset of the columns in ${\hyperref[ADef]{\bs{{\rm A}}'}}$ is linearly independent. \begin{myLemma} \label{basisLem} Let ${\hyperref[ADef]{\bs{{\rm A}}'}}$ be {\hyperref[mldDef]{minimally linearly dependent}}. Then for {\hyperref[aeDef]{{\rm a.e.}}}\ ${\hyperref[sstarDef]{S^\star}}$, ${\hyperref[nOfDef]{n}}({\hyperref[ADef]{\bs{{\rm A}}'}}) = {\hyperref[mOfDef]{m}}({\hyperref[ADef]{\bs{{\rm A}}'}}) - {\hyperref[rDef]{r}} + 1$. \end{myLemma} \begin{proof} Let \phantomsection\label{AdpDef}${\hyperref[ADef]{\bs{{\rm A}}'}}=[ \ {\hyperref[AdpDef]{\bs{{\rm A}}''}} \ | \ {\hyperref[aiDef]{\bs{{\rm a}}_i}} \ ]$ be {\hyperref[mldDef]{minimally linearly dependent}}. Let \phantomsection\label{mDef}${\hyperref[mDef]{m}}={\hyperref[mOfDef]{m}}({\hyperref[AdpDef]{\bs{{\rm A}}''}})$, \phantomsection\label{nDef}${\hyperref[nDef]{n}} = {\hyperref[nOfDef]{n}}({\hyperref[AdpDef]{\bs{{\rm A}}''}})$, and \phantomsection\label{liDef}${\hyperref[liDef]{\aleph}}={\hyperref[liOfDef]{\aleph}}({\hyperref[AdpDef]{\bs{{\rm A}}''}})$. Define \phantomsection\label{bbDef}$\hyperref[bbDef]{\bs{\beta}} \in \mathbb{R}^{\fix{{\hyperref[nDef]{n}}}}$ such that \begin{align} \label{aLdOnAEq} {\hyperref[AdpDef]{\bs{{\rm A}}''}} \hyperref[bbDef]{\bs{\beta}} \ = \ {\hyperref[aiDef]{\bs{{\rm a}}_i}} \ . \end{align} Note that because ${\hyperref[ADef]{\bs{{\rm A}}'}}$ is {\hyperref[mldDef]{minimally linearly dependent}}, all entries in $\hyperref[bbDef]{\bs{\beta}}$ are nonzero. Since the columns of ${\hyperref[AdpDef]{\bs{{\rm A}}''}}$ are linearly independent, ${\hyperref[nDef]{n}}={\hyperref[liDef]{\aleph}}$. Thus, by {\hyperref[liLem]{Lemma \ref{liLem}}}, ${\hyperref[nDef]{n}} \leq {\hyperref[mDef]{m}}-{\hyperref[rDef]{r}}$. We want to show that ${\hyperref[nDef]{n}}={\hyperref[mDef]{m}}-{\hyperref[rDef]{r}}$, so suppose for contradiction that ${\hyperref[nDef]{n}}<{\hyperref[mDef]{m}}-{\hyperref[rDef]{r}}$. We can assume without loss of generality that ${\hyperref[AdpDef]{\bs{{\rm A}}''}}$ has all its zero rows (if any) in the first positions. In that case, since ${\hyperref[ADef]{\bs{{\rm A}}'}}$ is {\hyperref[mldDef]{minimally linearly dependent}}, it follows that the nonzero entries of ${\hyperref[aiDef]{\bs{{\rm a}}_i}}$ cannot be in the corresponding rows. Thus, without loss of generality, assume that ${\hyperref[aiDef]{\bs{{\rm a}}_i}}$ has its first ${\hyperref[rDef]{r}}$ nonzero entries in the first ${\hyperref[rDef]{r}}$ nonzero rows of ${\hyperref[AdpDef]{\bs{{\rm A}}''}}$, and that the last nonzero entry of ${\hyperref[aiDef]{\bs{{\rm a}}_i}}$ is $1$ (i.e., rescale ${\hyperref[aiDef]{\bs{{\rm a}}_i}}$ if needed), and is located in the last row. Let \phantomsection\label{hataiDef}${\hyperref[hataiDef]{\bs{\hat{{\rm a}}}_i}} \in \mathbb{R}^{\fix{{\hyperref[rDef]{r}}}}$ denote the vector with the first nonzero entries of ${\hyperref[aiDef]{\bs{{\rm a}}_i}}$, such that we can write: \begin{align} \label{firstDecompositionAEq} \left[\begin{array}{c|c} {\hyperref[AdpDef]{\bs{{\rm A}}''}} & {\hyperref[aiDef]{\bs{{\rm a}}_i}} \end{array}\right] =\left[\begin{matrix} \\ \\ \\ \\ \\ \\ \\ \\ \end{matrix}\right. \underbrace{\begin{array}{c|} \hspace{.5cm} \Scale[1.5]{\bs{0}} \hspace{.5cm} \\ \hline \\ \Scale[1.5]{\bs{\hyperref[firstDecompositionAEq]{\bs{{\rm C}}}}} \\ \\ \hline \\ \multirow{2}{*}{\Scale[1.5]{\bs{\hyperref[firstDecompositionAEq]{\bs{{\rm B}}}}}} \\ \\ \\ \end{array}}_{\fix{{\hyperref[nDef]{n}}}{}} \underbrace{ \begin{array}{c} \vspace{.05cm} \Scale[1]{\bs{0}} \\ \hline \\ \Scale[1]{\bs{{\hyperref[hataiDef]{\bs{\hat{{\rm a}}}_i}}}} \\ \vspace{.05cm} \\ \hline \\ \Scale[1]{\bs{0}} \\ \\ \hline 1 \\ \end{array}}_{1} \left.\begin{matrix} \\ \\ \\ \\ \\ \\ \\ \\ \end{matrix}\right] \begin{matrix} \left. \begin{matrix} \\ \end{matrix} \right\} {\hyperref[dDef]{d}}-{\hyperref[mDef]{m}} \hspace{.6cm} \\ \left. \begin{matrix} \\ \\ \\ \end{matrix} \right\} {\hyperref[rDef]{r}} \hspace{1.2cm} \\ \left. \begin{matrix} \\ \\ \\ \end{matrix} \right\} {\hyperref[mDef]{m}}-{\hyperref[rDef]{r}}-1 \\ \left. \begin{matrix} \\ \end{matrix} \right\} 1, \hspace{1.1cm} \end{matrix} \end{align} where $\hyperref[firstDecompositionAEq]{\bs{{\rm C}}}$ and $\hyperref[firstDecompositionAEq]{\bs{{\rm B}}}$ are submatrices used to denote the blocks of ${\hyperref[AdpDef]{\bs{{\rm A}}''}}$ corresponding to the partition of ${\hyperref[aiDef]{\bs{{\rm a}}_i}}$. The columns of $\hyperref[firstDecompositionAEq]{\bs{{\rm B}}}$ are linearly independent. To see this, suppose for contradiction that they are not. This means that there exists some nonzero $\bs{\gamma} \in \mathbb{R}^{\fix{{\hyperref[nDef]{n}}}}$, such that $\hyperref[firstDecompositionAEq]{\bs{{\rm B}}} \bs{\gamma} = 0$. Let $\bs{c}={\hyperref[AdpDef]{\bs{{\rm A}}''}} \bs{\gamma} $ and note that only the ${\hyperref[rDef]{r}}$ rows in $\bs{c}$ corresponding to the block $\hyperref[firstDecompositionAEq]{\bs{{\rm C}}}$ may be nonzero. Let ${\hyperref[oDef]{\bs{\upsilon}}}$ denote the binary vector of these nonzero entries. Since ${\hyperref[sstarDef]{S^\star}}$ is orthogonal to every column of ${\hyperref[AdpDef]{\bs{{\rm A}}''}}$ and $\bs{c}$ is a linear combination of the columns in ${\hyperref[AdpDef]{\bs{{\rm A}}''}}$, it follows that ${\hyperref[cdotoDef]{S^\star_{\bs{\upsilon}}}} \subset \ker \bs{c}^\mathsf{T}_{\fix{{\hyperref[oDef]{\bs{\upsilon}}}}}$. This implies that $\dim {\hyperref[cdotoDef]{S^\star_{\bs{\upsilon}}}} \leq \dim \ker \bs{c}^\mathsf{T}_{\fix{{\hyperref[oDef]{\bs{\upsilon}}}}} = \|{\hyperref[oDef]{\bs{\upsilon}}}\|_1-1$. As in the proof of {\hyperref[aEntriesLem]{Lemma \ref{aEntriesLem}}}, this implies that the columns of $\hyperref[firstDecompositionAEq]{\bs{{\rm B}}}$ are linearly dependent only in a set of measure zero. Going back to \eqref{firstDecompositionAEq}, since the ${\hyperref[nDef]{n}}$ columns of $\hyperref[firstDecompositionAEq]{\bs{{\rm B}}}$ are linearly independent and because we are assuming that ${\hyperref[nDef]{n}}<{\hyperref[mDef]{m}}-{\hyperref[rDef]{r}}$, it follows that $\hyperref[firstDecompositionAEq]{\bs{{\rm B}}}$ has ${\hyperref[nDef]{n}}$ linearly independent rows. Let \phantomsection\label{BoneDef}$\hyperref[BoneDef]{\bs{{\rm B}}_1}$ denote the ${\hyperref[nDef]{n}} \times {\hyperref[nDef]{n}}$ block of $\hyperref[firstDecompositionAEq]{\bs{{\rm B}}}$ that contains ${\hyperref[nDef]{n}}$ linearly independent rows, and \phantomsection\label{BtwoDef}$\hyperref[BtwoDef]{\bs{{\rm B}}_2}$ the $({\hyperref[mDef]{m}}-{\hyperref[nDef]{n}}-{\hyperref[rDef]{r}}) \times {\hyperref[nDef]{n}}$ remaining block of $\hyperref[firstDecompositionAEq]{\bs{{\rm B}}}$. Notice that the row of $\hyperref[firstDecompositionAEq]{\bs{{\rm B}}}$ corresponding to the $1$ in ${\hyperref[aiDef]{\bs{{\rm a}}_i}}$ must belong to $\hyperref[BoneDef]{\bs{{\rm B}}_1}$, since otherwise, we have that $\hyperref[BoneDef]{\bs{{\rm B}}_1} \hyperref[bbDef]{\bs{\beta}} =0$, with $\hyperref[bbDef]{\bs{\beta}}$ as in \eqref{aLdOnAEq}, which implies that $\hyperref[BoneDef]{\bs{{\rm B}}_1}$ is rank deficient, in contradiction to its construction. We can further assume without loss of generality that the first nonzero entry of every column of $\hyperref[firstDecompositionAEq]{\bs{{\rm B}}}$ is $1$ (otherwise we may just rescale each column), and that these nonzero entries are in the first columns (otherwise we may just permute the columns accordingly). We will also let \phantomsection\label{BtwoTildeDef}$\hyperref[BtwoTildeDef]{\widetilde{\bs{{\rm B}}}_2}$ denote all but the first row of $\hyperref[BtwoDef]{\bs{{\rm B}}_2}$. Thus, our matrix is organized as \begin{align} \small \label{AconstructionEq} \left[\begin{array}{c|c} {\hyperref[AdpDef]{\bs{{\rm A}}''}} & {\hyperref[aiDef]{\bs{{\rm a}}_i}} \end{array}\right] = \begin{matrix} \begin{matrix} \\\vspace{.1cm} \\ \\ \\ \\ \end{matrix} \\ \left. \begin{matrix} \\ \\ \vspace{.3cm} \\ \\ \end{matrix} \hyperref[BtwoDef]{\bs{{\rm B}}_2}\right\{ \\ \begin{matrix} \\ \\ \\ \\ \end{matrix} \end{matrix} \left[\begin{array}{cc|c} \multicolumn{2}{c|}{\hspace{.75cm} \Scale[1.5]{\bs{0}} \hspace{.75cm}} & \bs{0}\\ \hline &&\\ \multicolumn{2}{c|}{\Scale[1.5]{\hyperref[firstDecompositionAEq]{\bs{{\rm C}}}}} & {\hyperref[hataiDef]{\bs{\hat{{\rm a}}}_i}} \\&&\\ \hline \multicolumn{1}{c|}{\hspace{.3cm} \Scale[1]{{\hyperref[oneDef]{\bs{{\rm 1}}}}} \hspace{.3cm}} & \hspace{.1cm} \Scale[1]{\bs{0}} & 0 \\ \hline &&\\ \multicolumn{2}{c|}{\Scale[1.5]{\hyperref[BtwoTildeDef]{\widetilde{\bs{{\rm B}}}_2}}} & \bs{0} \\&&\\ \hline &&\multirow{2}{*}{\Scale[1]{\bs{0}}} \\ \multicolumn{2}{c|}{\Scale[1.5]{\hyperref[BoneDef]{\bs{{\rm B}}_1}}} & \\ \cline{3-3}&&1\\ \end{array} \right] \begin{matrix} \left. \begin{matrix} \\ \end{matrix} \right\} {\hyperref[dDef]{d}}-{\hyperref[mDef]{m}} \hspace{.5cm} \\ \left. \begin{matrix} \\ \\ \vspace{.1cm} \\ \end{matrix} \right\} {\hyperref[rDef]{r}} \hspace{1.1cm} \\ \left. \begin{matrix} \\ \end{matrix} \right\} 1 \hspace{1.1cm} \\ \left. \begin{matrix} \\ \\ \vspace{.2cm} \\ \end{matrix} \right\} \begin{array}{l}{\hyperref[mDef]{m}}-{\hyperref[nDef]{n}} \\ -{\hyperref[rDef]{r}}-1 \\ \geq 0\end{array} \\ \left. \begin{matrix} \\ \\ \end{matrix} \right\} {\hyperref[nDef]{n}}-1 \hspace{.5cm} \\ \left. \begin{matrix} \\ \end{matrix} \right\} 1. \hspace{0.9cm} \end{matrix} \normalsize \end{align} Now \eqref{aLdOnAEq} implies $\hyperref[BoneDef]{\bs{{\rm B}}_1} \hyperref[bbDef]{\bs{\beta}} = [ \hspace{.1cm} \bs{0} \hspace{.1cm} | \hspace{.1cm} 1 \hspace{.1cm}]^\mathsf{T}$, and since $\hyperref[BoneDef]{\bs{{\rm B}}_1}$ is full rank, we may write \begin{align*} \hyperref[bbDef]{\bs{\beta}} \ = \ \hyperref[BoneDef]{\bs{{\rm B}}_1^{-1}} \left[\begin{matrix} \bs{0} \\ 1 \end{matrix}\right], \end{align*} i.e., $\hyperref[bbDef]{\bs{\beta}}$ is the the last column of the inverse of $\hyperref[BoneDef]{\bs{{\rm B}}_1}$, which is a rational function in the elements of $\hyperref[BoneDef]{\bs{{\rm B}}_1}$. Next, let us look back at \eqref{aLdOnAEq}. If ${\hyperref[nDef]{n}}<{\hyperref[mDef]{m}}-{\hyperref[rDef]{r}}$, then using the additional row $[ \hspace{.1cm} {\hyperref[oneDef]{\bs{{\rm 1}}}} \hspace{.1cm} | \hspace{.1cm} \bs{0} \hspace{.1cm} ]$ of \eqref{AconstructionEq} (which does not appear if ${\hyperref[mDef]{m}}={\hyperref[nDef]{n}}+{\hyperref[rDef]{r}}$) we obtain $[ \hspace{.1cm} {\hyperref[oneDef]{\bs{{\rm 1}}}} \hspace{.1cm} | \hspace{.1cm} \bs{0} \hspace{.1cm}] \hyperref[bbDef]{\bs{\beta}} = 0$. Recall that all the entries of $\hyperref[bbDef]{\bs{\beta}}$ are nonzero. Thus, the last equation defines the following nonzero rational function in the elements of $\hyperref[BoneDef]{\bs{{\rm B}}_1}$: \begin{align} \label{polyEq} \left[\begin{array}{c|r} {\hyperref[oneDef]{\bs{{\rm 1}}}} & \bs{0} \end{array}\right] \hyperref[BoneDef]{\bs{{\rm B}}_1^{-1}}\left[\begin{matrix} \bs{0} \\ 1 \end{matrix}\right] \ = \ 0 . \end{align} Equivalently, \eqref{polyEq} is a polynomial equation in the elements of $\hyperref[BoneDef]{\bs{{\rm B}}_1}$, which we will denote as ${\hyperref[polyEq]{f}}(\hyperref[BoneDef]{\bs{{\rm B}}_1})=0$. Next note that for {\hyperref[aeDef]{{\rm a.e.}}}\ ${\hyperref[sstarDef]{S^\star}}$, we can write \phantomsection\label{AAstarDef}${\hyperref[sstarDef]{S^\star}}=\ker {\hyperref[AAstarDef]{\bs{{\rm A}}^\star}}{}^\mathsf{T}$ for a unique ${\hyperref[AADef]{\bs{{\rm A}}}}^\star \in \mathbb{R}^{\fix{{\hyperref[dDef]{d}}} \times (\fix{{\hyperref[dDef]{d}}}-\fix{{\hyperref[rDef]{r}}})}$ in column echelon form\footnote{Certain ${\hyperref[sstarDef]{S^\star}}$ may not admit this representation, e.g., if ${\hyperref[sstarDef]{S^\star}}$ is orthogonal to certain canonical coordinates, which, as discussed in {\hyperref[aEntriesLem]{Lemma \ref{aEntriesLem}}}, is not the case for almost every ${\hyperref[sstarDef]{S^\star}}$ in $\hyperref[GrDef]{{\rm Gr}}({\hyperref[rDef]{r}},\mathbb{R}^{\fix{{\hyperref[dDef]{d}}}})$.}: \begin{align} \label{AAstarEq} {\hyperref[AAstarDef]{\bs{{\rm A}}^\star}} \ = \ \left[ \begin{array}{c} \\ \Scale[2]{{\hyperref[IDef]{\bs{{\rm I}}}}} \\ \\ \hline \hspace{.4cm} \Scale[1.5]{\bs{{\hyperref[AAstarEq]{\bs{{\rm D}}^\star}}}} \hspace{.2cm} \\ \end{array}\right] \begin{matrix} \left. \begin{matrix} \\ \\ \\ \end{matrix} \right\} {\hyperref[dDef]{d}}-{\hyperref[rDef]{r}} \\ \left. \begin{matrix} \\ \end{matrix} \right\} {\hyperref[rDef]{r}} \hspace{.5cm}. \end{matrix} \end{align} On the other hand every ${\hyperref[AAstarEq]{\bs{{\rm D}}^\star}} \in \mathbb{R}^{\fix{{\hyperref[rDef]{r}}} \times (\fix{{\hyperref[dDef]{d}}}-\fix{{\hyperref[rDef]{r}}})}$ defines a unique ${\hyperref[rDef]{r}}$-dimensional subspace of $\mathbb{R}^{\fix{{\hyperref[dDef]{d}}}}$, via \eqref{AAstarEq}. Thus, we have a bijection between $\mathbb{R}^{\fix{{\hyperref[rDef]{r}}} \times (\fix{{\hyperref[dDef]{d}}}-\fix{{\hyperref[rDef]{r}}})}$ and a dense open subset of $\hyperref[GrDef]{{\rm Gr}}({\hyperref[rDef]{r}},\mathbb{R}^{\fix{{\hyperref[dDef]{d}}}})$. Since the columns of ${\hyperref[AdpDef]{\bs{{\rm A}}''}}$ must be linear combinations of the columns of ${\hyperref[AAstarDef]{\bs{{\rm A}}^\star}}$, the elements of $\hyperref[BoneDef]{\bs{{\rm B}}_1}$ are linear functions in the entries of ${\hyperref[AAstarEq]{\bs{{\rm D}}^\star}}$. Therefore, we can express ${\hyperref[polyEq]{f}}(\hyperref[BoneDef]{\bs{{\rm B}}_1})$ as a nonzero polynomial function $g$ in the entries of ${\hyperref[AAstarEq]{\bs{{\rm D}}^\star}}$ and rewrite \eqref{polyEq} as $g({\hyperref[AAstarEq]{\bs{{\rm D}}^\star}})=0$. But we know that $g({\hyperref[AAstarEq]{\bs{{\rm D}}^\star}}) \neq 0$ for almost every ${\hyperref[AAstarEq]{\bs{{\rm D}}^\star}} \in \mathbb{R}^{\fix{{\hyperref[rDef]{r}}} \times (\fix{{\hyperref[dDef]{d}}}-\fix{{\hyperref[rDef]{r}}})}$, and hence for almost every ${\hyperref[sstarDef]{S^\star}} \in \hyperref[GrDef]{{\rm Gr}}({\hyperref[rDef]{r}},\mathbb{R}^{\fix{{\hyperref[dDef]{d}}}})$. We conclude that almost every subspace in $\hyperref[GrDef]{{\rm Gr}}({\hyperref[rDef]{r}},\mathbb{R}^{\fix{{\hyperref[dDef]{d}}}})$ will not satisfiy \eqref{polyEq}, and thus ${\hyperref[nDef]{n}} = {\hyperref[mDef]{m}} - {\hyperref[rDef]{r}}$. \end{proof} We are now ready to present the proofs of {\hyperref[independenceLem]{Lemma \ref{independenceLem}}}\ and {\hyperref[identifiabilityThm]{Theorem \ref{identifiabilityThm}}}. \begin{proof}({\hyperref[independenceLem]{Lemma \ref{independenceLem}}}) \begin{itemize} \item[($\Rightarrow$)] Suppose ${\hyperref[ADef]{\bs{{\rm A}}'}}$ is {\hyperref[mldDef]{minimally linearly dependent}}. By {\hyperref[basisLem]{Lemma \ref{basisLem}}}, ${\hyperref[nOfDef]{n}}({\hyperref[ADef]{\bs{{\rm A}}'}})={\hyperref[mOfDef]{m}}({\hyperref[ADef]{\bs{{\rm A}}'}})-{\hyperref[rDef]{r}}+1>{\hyperref[mOfDef]{m}}({\hyperref[ADef]{\bs{{\rm A}}'}})-{\hyperref[rDef]{r}}$, and we have the first implication. \item[($\Leftarrow$)] Suppose there exists an ${\hyperref[ADef]{\bs{{\rm A}}'}}$ with ${\hyperref[nOfDef]{n}}({\hyperref[ADef]{\bs{{\rm A}}'}})>{\hyperref[mOfDef]{m}}({\hyperref[ADef]{\bs{{\rm A}}'}})-{\hyperref[rDef]{r}}$. By {\hyperref[liLem]{Lemma \ref{liLem}}}, ${\hyperref[nOfDef]{n}}({\hyperref[ADef]{\bs{{\rm A}}'}})>{\hyperref[liOfDef]{\aleph}}({\hyperref[ADef]{\bs{{\rm A}}'}})$, which implies the columns in ${\hyperref[ADef]{\bs{{\rm A}}'}}$, and hence ${\hyperref[AADef]{\bs{{\rm A}}}}$, are linearly dependent. \end{itemize} \end{proof} \begin{proof}({\hyperref[identifiabilityThm]{Theorem \ref{identifiabilityThm}}}) {\hyperref[aEntriesLem]{Lemma \ref{aEntriesLem}}}\ shows that for {\hyperref[aeDef]{{\rm a.e.}}}\ ${\hyperref[sstarDef]{S^\star}}$, the $({\hyperref[jDef]{j}},{\hyperref[iDef]{i}})^{th}$ entry of ${\hyperref[AADef]{\bs{{\rm A}}}}$ is nonzero if and only if the $({\hyperref[jDef]{j}},{\hyperref[iDef]{i}})^{th}$ entry of ${\hyperref[OODef]{\bs{\Omega}}}$ is nonzero. \begin{itemize} \item[($\Rightarrow$)] Suppose there exists an ${\hyperref[ODef]{\bs{\Omega}'}}$ such that ${\hyperref[mOfDef]{m}}({\hyperref[ODef]{\bs{\Omega}'}}) < {\hyperref[nOfDef]{n}}({\hyperref[ODef]{\bs{\Omega}'}})+{\hyperref[rDef]{r}}$. Then ${\hyperref[mOfDef]{m}}({\hyperref[ADef]{\bs{{\rm A}}'}}) < {\hyperref[nOfDef]{n}}({\hyperref[ADef]{\bs{{\rm A}}'}})+{\hyperref[rDef]{r}}$ for some ${\hyperref[ADef]{\bs{{\rm A}}'}}$. {\hyperref[independenceLem]{Lemma \ref{independenceLem}}}\ implies that the columns of ${\hyperref[ADef]{\bs{{\rm A}}'}}$, and hence ${\hyperref[AADef]{\bs{{\rm A}}}}$, are linearly dependent. This implies $\dim \ker{\hyperref[AADef]{\bs{{\rm A}}}}{}^\mathsf{T} > {\hyperref[rDef]{r}}$. \item[($\Leftarrow$)] Suppose every ${\hyperref[ODef]{\bs{\Omega}'}}$ satisfies ${\hyperref[mOfDef]{m}}({\hyperref[ODef]{\bs{\Omega}'}}) \geq {\hyperref[nOfDef]{n}}({\hyperref[ODef]{\bs{\Omega}'}})+{\hyperref[rDef]{r}}$. Then ${\hyperref[mOfDef]{m}}({\hyperref[ADef]{\bs{{\rm A}}'}}) \geq {\hyperref[nOfDef]{n}}({\hyperref[ADef]{\bs{{\rm A}}'}})+{\hyperref[rDef]{r}}$ for every ${\hyperref[ADef]{\bs{{\rm A}}'}}$, including ${\hyperref[AADef]{\bs{{\rm A}}}}$. Therefore, by {\hyperref[independenceLem]{Lemma \ref{independenceLem}}}, the ${\hyperref[dDef]{d}}-{\hyperref[rDef]{r}}$ columns in ${\hyperref[AADef]{\bs{{\rm A}}}}$ are linearly independent, hence $\dim \ker {\hyperref[AADef]{\bs{{\rm A}}}}{}^\mathsf{T}={\hyperref[rDef]{r}}$. \end{itemize} \end{proof} \vspace{.3cm} \section{Implications for low-rank matrix completion} \label{LRMCSec} \vspace{.3cm} Subspace identifiability is closely related to the \phantomsection\label{LRMCDef}low-rank matrix completion ({\hyperref[LRMCDef]{LRMC}}) problem \cite{candes}: given a subset of entries in a rank-${\hyperref[rDef]{r}}$ matrix, exactly recover {\em all} of the missing entries. This requires, implicitly, idenficiation of the subspace spanned by the complete columns of the matrix. We use this section to present the implications of our results for {\hyperref[LRMCDef]{LRMC}}. Let \phantomsection\label{XDef}${\hyperref[XDef]{\bs{{\rm X}}}}$ be a ${\hyperref[dDef]{d}} \times {\hyperref[NDef]{N}}$, rank-${\hyperref[rDef]{r}}$ matrix and assume that \begin{itemize} \phantomsection\label{AthreeAssDef} \item[{\hyperref[AthreeAssDef]{\textbf{A3}}}] The columns of ${\hyperref[XDef]{\bs{{\rm X}}}}$ are drawn independently according to \phantomsection\label{nuuDef}${\hyperref[nuuDef]{\nu}}$, an absolutely continuous distribution with respect to the Lebesgue measure on ${\hyperref[sstarDef]{S^\star}}$. \end{itemize} Let \phantomsection\label{XODef}${\hyperref[XODef]{\bs{{\rm X}}_{\bs{\Omega}}}}$ be the incomplete version of ${\hyperref[XDef]{\bs{{\rm X}}}}$, observed only in the nonzero positions of ${\hyperref[OODef]{\bs{\Omega}}}$. \subsection*{Necessary and sufficient conditions for {\hyperref[LRMCDef]{LRMC}}} To relate the {\hyperref[LRMCDef]{LRMC}}\ problem to our main results, define \phantomsection\label{NtildeDef}${\hyperref[NtildeDef]{\widetilde{N}}}$ as the number of {\em distinct} columns (sampling patterns) in ${\hyperref[OODef]{\bs{\Omega}}}$, and let \phantomsection\label{OTildeDef}${\hyperref[OTildeDef]{\widetilde{\bs{\Omega}}}}$ denote a ${\hyperref[dDef]{d}} \times {\hyperref[NtildeDef]{\widetilde{N}}}$ matrix composed of these columns. \begin{myCorollary} \label{LRMCnecCor} If ${\hyperref[OTildeDef]{\widetilde{\bs{\Omega}}}}$ does not contain a ${\hyperref[dDef]{d}} \times ({\hyperref[dDef]{d}}-{\hyperref[rDef]{r}})$ submatrix satisfying the conditions of {\hyperref[identifiabilityThm]{Theorem \ref{identifiabilityThm}}}, then ${\hyperref[XDef]{\bs{{\rm X}}}}$ cannot be uniquely recovered from ${\hyperref[XODef]{\bs{{\rm X}}_{\bs{\Omega}}}}$. \end{myCorollary} Since ${\hyperref[XDef]{\bs{{\rm X}}}}$ is rank-${\hyperref[rDef]{r}}$, a column with fewer than ${\hyperref[rDef]{r}}$ observed entries cannot be completed (in general). We will thus assume without loss of generality the following relaxation of {\hyperref[AoneAssDef]{\textbf{A1}}}: \begin{itemize} \phantomsection\label{AonepAssDef} \item[{\hyperref[AonepAssDef]{\textbf{A1'}}}] ${\hyperref[OODef]{\bs{\Omega}}}$ has at least ${\hyperref[rDef]{r}}$ nonzero entries per column. \end{itemize} \begin{myCorollary} \label{LRMCsuffCor} Let {\hyperref[AonepAssDef]{\textbf{A1'}}}\ and {\hyperref[AthreeAssDef]{\textbf{A3}}}\ hold. Suppose ${\hyperref[OTildeDef]{\widetilde{\bs{\Omega}}}}$ contains a ${\hyperref[dDef]{d}} \times ({\hyperref[dDef]{d}}-{\hyperref[rDef]{r}})$ submatrix satisfying the conditions of {\hyperref[identifiabilityThm]{Theorem \ref{identifiabilityThm}}}, and that for every column ${\hyperref[oiDef]{\bs{\omega}_i}}$ in this submatrix, at least ${\hyperref[rDef]{r}}$ columns in ${\hyperref[XODef]{\bs{{\rm X}}_{\bs{\Omega}}}}$ are observed at the nonzero locations of ${\hyperref[oiDef]{\bs{\omega}_i}}$. Then for {\hyperref[aeDef]{{\rm a.e.}}}\ ${\hyperref[sstarDef]{S^\star}}$, and almost surely with respect to ${\hyperref[nuuDef]{\nu}}$, ${\hyperref[XDef]{\bs{{\rm X}}}}$ can be uniquely recovered from ${\hyperref[XODef]{\bs{{\rm X}}_{\bs{\Omega}}}}$. \end{myCorollary} Proofs of these results are given in the {\hyperref[LRMCnecApx]{appendix}}. The intuition behind {\hyperref[LRMCnecCor]{Corollary \ref{LRMCnecCor}}}\ is simply that identifying a subspace from its projections onto sets of canonical coordinates is {\em easier } than {\hyperref[LRMCDef]{LRMC}}, and so the necessary condition of {\hyperref[identifiabilityThm]{Theorem \ref{identifiabilityThm}}} \ is also necessary for {\hyperref[LRMCDef]{LRMC}}. {\hyperref[LRMCsuffCor]{Corollary \ref{LRMCsuffCor}}}\ follows from the fact that ${\hyperref[sstarDef]{S^\star}}$ (or its projections) can be determined from ${\hyperref[rDef]{r}}$ or more observations drawn from ${\hyperref[nuuDef]{\nu}}$. \subsection*{Validating {\hyperref[LRMCDef]{LRMC}}} Under certain assumptions on the subset of observed entries (e.g., random sampling) and ${\hyperref[sstarDef]{S^\star}}$ (e.g., incoherence), existing methods, for example nuclear norm minimization \cite{candes}, succeed {\em with high probability} in completing the matrix exactly and thus identifying ${\hyperref[sstarDef]{S^\star}}$. These assumptions are sufficient, but not necessary, and are sometimes unverifiable or unjustified in practice. Therefore, the result of an {\hyperref[LRMCDef]{LRMC}}\ algorithm can be suspect. Simply finding a low-rank matrix that agrees with the observed data does not guarantee that it is the correct completion. It is possible that there exist other ${\hyperref[rDef]{r}}$-dimensional subspaces different from ${\hyperref[sstarDef]{S^\star}}$ that agree with the observed entries. \begin{myExample} Suppose we run an {\hyperref[LRMCDef]{LRMC}}\ algorithm on a matrix observed on the support of ${\hyperref[OODef]{\bs{\Omega}}}$, with ${\hyperref[OODef]{\bs{\Omega}}}$ and ${\hyperref[sstarDef]{S^\star}}$ as in {\hyperref[introEg]{Example \ref{introEg}}} \ in Section~\ref{modelSec}. Suppose that the algorithm produces a completion with columns from ${\hyperref[sDef]{S}}={\rm span}[1 \ \ 2 \ \ 3 \ \ 5 \ \ 5]^\mathsf{T}$ instead of ${\hyperref[sstarDef]{S^\star}}$. It is clear that the residual of the projection of any vector from ${\hyperref[soiDef]{S^\star_{\bs{\omega}_i}}}$ onto ${\hyperref[soiDef]{S_{\bs{\omega}_i}}}$ will be zero, despite the fact that ${\hyperref[sDef]{S}} \neq {\hyperref[sstarDef]{S^\star}}$. \end{myExample} In other words, if the residuals are nonzero, we can discard an incorrect solution, but if the residuals are zero, we cannot validate whether our solution is correct or not. \vspace{.3cm} \begingroup \leftskip1.5em \rightskip\leftskip \noindent {\em {\hyperref[fitsCor]{Corollary \ref{fitsCor}}}, below, allows one to drop the sampling and incoherence assumptions, and validate the result of {\em any} {\hyperref[LRMCDef]{LRMC}}\ algorithm deterministically.} \par \endgroup \vspace{.3cm} Let \phantomsection\label{xDef}${\hyperref[xDef]{\bs{{\rm x}}}}_i$ denote the ${\hyperref[iDef]{i}}^{th}$ column of ${\hyperref[XDef]{\bs{{\rm X}}}}$, and \phantomsection\label{xoiDef}${\hyperref[xoiDef]{\bs{{\rm x}}_{\bs{\omega}_i}}}$ be the restriction of ${\hyperref[xDef]{\bs{{\rm x}}}}_i$ to the nonzero coordinates of ${\hyperref[oiDef]{\bs{\omega}_i}}$. We say that a subspace ${\hyperref[sDef]{S}}$ \phantomsection\label{fitsDef}{\hyperref[fitsDef]{fits}}\ ${\hyperref[XODef]{\bs{{\rm X}}_{\bs{\Omega}}}}$ if ${\hyperref[xoiDef]{\bs{{\rm x}}_{\bs{\omega}_i}}} \in {\hyperref[soiDef]{S_{\bs{\omega}_i}}}$ for every ${\hyperref[iDef]{i}}$. \begin{myCorollary} \label{fitsCor} Let {\hyperref[AthreeAssDef]{\textbf{A3}}}\ hold, and suppose ${\hyperref[XODef]{\bs{{\rm X}}_{\bs{\Omega}}}}$ contains two disjoint sets of columns, \phantomsection\label{XOoneDef}${\hyperref[XOoneDef]{\bs{{\rm X}}_{\bs{\Omega}_1}}}$ and \phantomsection\label{XOtwoDef}${\hyperref[XOtwoDef]{\bs{{\rm X}}_{\bs{\Omega}_2}}}$, such that \phantomsection\label{OOtwoDef}${\hyperref[XOtwoDef]{\bs{\Omega}_2}}$ is a ${\hyperref[dDef]{d}} \times ({\hyperref[dDef]{d}}-{\hyperref[rDef]{r}})$ matrix satisfying the conditions of {\hyperref[identifiabilityThm]{Theorem \ref{identifiabilityThm}}}. Let ${\hyperref[sDef]{S}}$ be the subspace spanned by the columns of a completion of ${\hyperref[XOoneDef]{\bs{{\rm X}}_{\bs{\Omega}_1}}}$. Then for {\hyperref[aeDef]{{\rm a.e.}}}\ ${\hyperref[sstarDef]{S^\star}}$, and almost surely with respect to ${\hyperref[nuuDef]{\nu}}$, ${\hyperref[sDef]{S}}$ {\hyperref[fitsDef]{fits}}\ ${\hyperref[XOtwoDef]{\bs{{\rm X}}_{\bs{\Omega}_2}}}$ if and only if ${\hyperref[sDef]{S}}={\hyperref[sstarDef]{S^\star}}$. \end{myCorollary} The proof of {\hyperref[fitsCor]{Corollary \ref{fitsCor}}}\ is given in the {\hyperref[fitsApx]{appendix}}. In words, {\hyperref[fitsCor]{Corollary \ref{fitsCor}}}\ states that if one runs an {\hyperref[LRMCDef]{LRMC}}\ algorithm on ${\hyperref[XOoneDef]{\bs{{\rm X}}_{\bs{\Omega}_1}}}$, then the uniqueness and correctness of the resulting low-rank completion can be verified by testing whether it agrees with the {\em validation} set ${\hyperref[XOtwoDef]{\bs{{\rm X}}_{\bs{\Omega}_2}}}$. \begin{myExample} Consider a $1000 \times 2000$ matrix ${\hyperref[XODef]{\bs{{\rm X}}_{\bs{\Omega}}}}$ with ${\hyperref[rDef]{r}}=30$ and ideal incoherence. In this case, the best sufficient conditions for {\hyperref[LRMCDef]{LRMC}}\ that we are aware of \cite{recht} require that all entries are observed. Simulations show that alternating minimization \cite{jain} can exactly complete such matrices when fewer than half of the entries are observed, and only using half of the columns. While previous theory for matrix completion gives no guarantees in scenarios like this, our new results do. To see this, split ${\hyperref[XODef]{\bs{{\rm X}}_{\bs{\Omega}}}}$ into two $1000 \times 1000$ submatrices ${\hyperref[XOoneDef]{\bs{{\rm X}}_{\bs{\Omega}_1}}}$ and ${\hyperref[XOtwoDef]{\bs{{\rm X}}_{\bs{\Omega}_2}}}$. Use nuclear norm, alternating minimization, or any {\hyperref[LRMCDef]{LRMC}}\ method, to find a completion of ${\hyperref[XOoneDef]{\bs{{\rm X}}_{\bs{\Omega}_1}}}$. {\hyperref[probabilityThm]{Theorem \ref{probabilityThm}}}\ can be used to show that the sampling of ${\hyperref[XOtwoDef]{\bs{{\rm X}}_{\bs{\Omega}_2}}}$ will satisfy the conditions of {\hyperref[identifiabilityThm]{Theorem \ref{identifiabilityThm}}}\ {\hyperref[whpDef]{w.h.p}}. even when only half the entries are observed randomly. We can then use {\hyperref[fitsCor]{Corollary \ref{fitsCor}}}\ to show that if ${\hyperref[XOtwoDef]{\bs{{\rm X}}_{\bs{\Omega}_2}}}$ is consistent with the completion of ${\hyperref[XOoneDef]{\bs{{\rm X}}_{\bs{\Omega}_1}}}$, then the completion is unique and correct. \end{myExample} \subsection*{Remarks} Observe that the necessary and sufficient conditions in Corollaries \ref{LRMCnecCor} and \ref{LRMCsuffCor} and the validation in {\hyperref[fitsCor]{Corollary \ref{fitsCor}}}\ do not require the incoherence assumptions typically needed in {\hyperref[LRMCDef]{LRMC}}\ results in order to guarantee correctness and uniqueness. Another advantage of results above is that they work for matrices of any rank, while standard {\hyperref[LRMCDef]{LRMC}}\ results only hold for ranks significantly smaller than the dimension ${\hyperref[dDef]{d}}$. Finally, the results above hold with probability $1$, as opposed to standard {\hyperref[LRMCDef]{LRMC}}\ statements, that hold {\hyperref[whpDef]{w.h.p}}. On the other hand, verifying whether ${\hyperref[XOtwoDef]{\bs{\Omega}_2}}$ meets the conditions of {\hyperref[identifiabilityThm]{Theorem \ref{identifiabilityThm}}}\ may be difficult. Nevertheless, if the entries in our data matrix are sampled randomly with rates comparable to standard conditions in {\hyperref[LRMCDef]{LRMC}}, we know by {\hyperref[probabilityThm]{Theorem \ref{probabilityThm}}}\ that {\hyperref[whpDef]{w.h.p}}. ${\hyperref[XOtwoDef]{\bs{\Omega}_2}}$ will satisfy such conditions. \section{Graphical interpretation of the problem} \label{graphSec} The problem of {\hyperref[LRMCDef]{LRMC}}\ has also been studied from the graph theory perspective. For example, it has been shown that graph connectivity is a necessary condition for completion \cite{kiraly}. Being subspace identifiability so tightly related to {\hyperref[LRMCDef]{LRMC}}, it comes as no surprise that there also exist graph conditions for subspace identifiability. In this section we draw some connections between subspace identifiability and graph theory that give insight on the conditions in {\hyperref[identifiabilityThm]{Theorem \ref{identifiabilityThm}}}. We use this interpretation to show that graph connectivity is a necessary yet insufficient condition for subspace identification. Define \phantomsection\label{GODef}${\hyperref[GODef]{\mathcal{G}(\bs{\Omega})}}$ as the bipartite graph with disjoint sets of {\em row} and {\em column} vertices, where there is an edge between row vertex ${\hyperref[jDef]{j}}$ and column vertex ${\hyperref[iDef]{i}}$ if the $({\hyperref[jDef]{j}},{\hyperref[iDef]{i}})^{th}$ entry of ${\hyperref[OODef]{\bs{\Omega}}}$ is nonzero. \begin{myExample} \label{graphEg} With ${\hyperref[dDef]{d}}=5$, ${\hyperref[rDef]{r}}=1$ and \begin{center} \begin{tikzpicture} \node [title] (c) at (-3,-1) { ${\hyperref[OODef]{\bs{\Omega}}} = \left[\begin{matrix} 1 & 0 & 0 & 0 \\ 1 & 1 & 0 & 0 \\ 0 & 1 & 1 & 0 \\ 0 & 0 & 1 & 1 \\ 0 & 0 & 0 & 1 \\ \end{matrix}\right]$}; \node [title] (c) at (-1,-1) {$\Rightarrow$}; \node [title] (c) at (1,-2.25) {${\hyperref[GODef]{\mathcal{G}(\bs{\Omega})}}$}; \node [title] (c) at (2,.5) {Columns}; \node [title] (r) at (0,.5) {Rows}; \node [label] (c1) at (2,-0.25){$1$}; \node [vertex] (c1) at (2,-0.25){}; \node [label] (c2) at (2,-.75) {$2$}; \node [vertex] (c2) at (2,-.75) {}; \node [label] (c3) at (2,-1.25) {$3$}; \node [vertex] (c3) at (2,-1.25) {}; \node [label] (c4) at (2,-1.75) {$4$}; \node [vertex] (c4) at (2,-1.75) {}; \node [label] (r1) at (0,0) {$1$}; \node [vertex] (r1) at (0,0) {}; \node [label] (r2) at (0,-.5) {$2$}; \node [vertex] (r2) at (0,-.5) {}; \node [label] (r3) at (0,-1) {$3$}; \node [vertex] (r3) at (0,-1) {}; \node [label] (r4) at (0,-1.5) {$4$}; \node [vertex] (r4) at (0,-1.5) {}; \node [label] (r5) at (0,-2) {$5$}; \node [vertex] (r5) at (0,-2) {}; \draw [font=\scriptstyle] (c1) edge (r1) (c1) edge (r2) (c2) edge (r2) (c2) edge (r3) (c3) edge (r3) (c3) edge (r4) (c4) edge (r4) (c4) edge (r5); \end{tikzpicture}. \end{center} \end{myExample} Recall that the \phantomsection\label{neighborhoodDef}{\em {\hyperref[neighborhoodDef]{neighborhood}}} of a set of vertices is the collection of all their adjacent vertices. \vspace{.3cm} \begingroup \leftskip1.5em \rightskip\leftskip \noindent {\em The graph theoretic interpretation of the condition on ${\hyperref[OODef]{\bs{\Omega}}}$ in {\hyperref[identifiabilityThm]{Theorem \ref{identifiabilityThm}}}\ is that every set of ${\hyperref[nOfDef]{n}}$ column vertices in ${\hyperref[GODef]{\mathcal{G}(\bs{\Omega})}}$ must have a {\hyperref[neighborhoodDef]{neighborhood}}\ of at least ${\hyperref[nOfDef]{n}}+{\hyperref[rDef]{r}}$ row vertices.} \par \endgroup \vspace{.3cm} \begin{myExample} One may verify that every set of ${\hyperref[nOfDef]{n}}$ column vertices in ${\hyperref[GODef]{\mathcal{G}(\bs{\Omega})}}$ from {\hyperref[graphEg]{Example \ref{graphEg}}}\ has a {\hyperref[neighborhoodDef]{neighborhood}}\ of at least ${\hyperref[nOfDef]{n}}+{\hyperref[rDef]{r}}$ row vertices. \pagebreak On the other hand, if we consider ${\hyperref[OODef]{\bs{\Omega}}}$ as in {\hyperref[introEg]{Example \ref{introEg}}}, the {\hyperref[neighborhoodDef]{neighborhood}}\ of the column vertices $\{1,2,3\}$ in ${\hyperref[GODef]{\mathcal{G}(\bs{\Omega})}}$ contains fewer than ${\hyperref[nOfDef]{n}}+{\hyperref[rDef]{r}}$ row vertices: \begin{center} \begin{tikzpicture} \node [title] (c) at (1,-2.25) {${\hyperref[GODef]{\mathcal{G}(\bs{\Omega})}}$}; \node [title] (c) at (2,.5) {Columns}; \node [title] (r) at (0,.5) {Rows}; \node [label] (c1) at (2,-0.25){$1$}; \node [vertex, line width=1.25pt] (c1) at (2,-0.25){}; \node [label] (c2) at (2,-.75) {$2$}; \node [vertex, line width=1.25pt] (c2) at (2,-.75) {}; \node [label] (c3) at (2,-1.25) {$3$}; \node [vertex, line width=1.25pt] (c3) at (2,-1.25) {}; \node [label] (c4) at (2,-1.75) {$4$}; \node [vertex] (c4) at (2,-1.75) {}; \node [label] (r1) at (0,0) {$1$}; \node [vertex, line width=1.25pt] (r1) at (0,0) {}; \node [label] (r2) at (0,-.5) {$2$}; \node [vertex, line width=1.25pt] (r2) at (0,-.5) {}; \node [label] (r3) at (0,-1) {$3$}; \node [vertex, line width=1.25pt] (r3) at (0,-1) {}; \node [label] (r4) at (0,-1.5) {$4$}; \node [vertex] (r4) at (0,-1.5) {}; \node [label] (r5) at (0,-2) {$5$}; \node [vertex] (r5) at (0,-2) {}; \draw [font=\scriptstyle] (c1) edge[line width=.75pt] (r1) (c1) edge[line width=.75pt] (r2) (c2) edge[line width=.75pt] (r2) (c2) edge[line width=.75pt] (r3) (c3) edge[line width=.75pt] (r1) (c3) edge[line width=.75pt] (r3) (c4) edge (r4) (c4) edge (r5); \end{tikzpicture}. \end{center} \end{myExample} With this interpretation of {\hyperref[identifiabilityThm]{Theorem \ref{identifiabilityThm}}}, we can extend terms and results from graph theory to our context. One example is the next corollary, which states that ${\hyperref[rDef]{r}}$-row-connectivity is a necessary but insufficient condition for subspace identifiability. We say ${\hyperref[GODef]{\mathcal{G}(\bs{\Omega})}}$ is {\em ${\hyperref[rDef]{r}}$-row-connected} if ${\hyperref[GODef]{\mathcal{G}(\bs{\Omega})}}$ remains a connected graph after removing any set of ${\hyperref[rDef]{r}}-1$ row vertices and all their adjacent edges. \begin{myCorollary} \label{rConnectedCor} For {\hyperref[aeDef]{{\rm a.e.}}}\ ${\hyperref[sstarDef]{S^\star}}$, $|{\hyperref[SSDef]{\mathscr{S}}}({\hyperref[sstarDef]{S^\star}},{\hyperref[OODef]{\bs{\Omega}}})|>1$ if ${\hyperref[GODef]{\mathcal{G}(\bs{\Omega})}}$ is not ${\hyperref[rDef]{r}}$-row-connected. The converse is only true for ${\hyperref[rDef]{r}}=1$. \end{myCorollary} {\hyperref[rConnectedCor]{Corollary \ref{rConnectedCor}}}\ is proved in the {\hyperref[rConnectedApx]{appendix}}. \section{Conclusions} \label{conclusionsSec} In this paper we determined when and only when can one identify a subspace from its projections onto subsets of the canonical coordinates. We show that the conditions for identifiability hold {\hyperref[whpDef]{w.h.p}}. under standard random sampling schemes, and that when these conditions are met, identifying the subspace becomes a trivial task. This gives new necessary and sufficient conditions for {\hyperref[LRMCDef]{LRMC}}, and allows one to verify whether the result of {\em any} {\hyperref[LRMCDef]{LRMC}}\ algorithm is unique and correct without prior incoherence or sampling assumptions.
\section{Introduction}\label{sec:intro} Galaxy gas-phase metallicities encode information about the history of gas accretion, star formation, and gaseous outflows. Measurements of metallicity combined with accumulated stellar mass and gas content therefore provide stringent constraints on the baryonic processes relevant to galaxy formation. Current observations show that essentially all galaxies have low gas-phase metallicities compared to estimated chemical yields from star formation, with greater discrepancy at lower stellar masses \citep[the mass-metallicity relation, e.g.,][]{Tremonti2004, Lequeux1979}. This is commonly attributed to outflows of metal-enriched gas driven by intense star formation, supported by observations that such feedback is ubiquitous among galaxies with high densities of star formation \citep{Heckman2001, Newman2012}. The {\it spatial} distribution of metallicity within a galaxy further constrains the baryonic assembly history, in particular the effects of outflows and dynamical evolution. Star-forming disk galaxies in the local universe have negative radial metallicity gradients (higher metallicity in the central regions; see \citealt{Sanchez2014} and \citealt{Vila-Costas1992} for large samples). This has been explained by models of chemical evolution in which galaxies grow {\it inside-out} such that outer regions have younger characteristic ages \citep[e.g.,][]{Nelson2012}. Gradients typically flatten at large radii indicating efficient mixing in those regions, possibly due to secular dynamical evolution or recycling of metal-enriched gas which was previously ejected in outflows \citep[so-called "galactic fountains"; e.g.][]{Werk2011, Bresolin2012}. Notably, galaxies which are undergoing mergers or strong interactions have significantly shallower metallicity gradients than isolated disks, which is understood numerically in terms of radial gas flows induced by tidal forces \citep[most notably metal-poor inflows;][]{Rupke2010a, Rupke2010b, Sanchez2014, Rich2012}. This work is concerned with using time evolution of metallicity gradients as a probe of galaxy formation. We seek to measure how gradients evolve, and to understand what processes drive that evolution. A variety of predictions have been made based on numerical models of inside-out growth, ranging from steeper to flatter to inverted (positive) gradients at higher redshifts \citep[e.g.,][]{Prantzos2000, Chiappini2001, Magrini2007, Fu2009}. Cosmological simulations have recently begun to address this issue, with several groups arguing that gradient evolution depends strongly on star formation feedback: stronger feedback is predicted to cause flatter gradients due to rapid gas recycling \citep{Pilkington2012, Gibson2013, Angles-Alcazar2014}. Meanwhile the observational results at high redshift have grown in number but comprise a variety of conclusions for the evolution of gradients. Data reaching $\lesssim1$ kpc resolution with adaptive optics (and up to $\sim$100 pc in the case of lensed galaxies) have revealed negative radial gradients with slopes often steeper than local descendants, suggesting that gradients flatten as galaxies grow with time \citep{Jones2010,Jones2013,Yuan2011,Swinbank2012}. Larger surveys with seeing-limited resolution of $\sim$5 kpc have reported mostly flat gradients with a significant fraction of positive slopes \citep{Cresci2010,Queyrel2012,Stott2014,Troncoso2014}. These have been interpreted as evidence for predicted "cold flows" of pristine gas \citep[e.g.,][]{Dekel2009,Keres2009,Faucher-Giguere2011}, but these results are in contrast with high-resolution measurements. The discrepancy in metallicity gradient measurements at high redshift must be addressed in order to reliably understand the role of gas and metal transport in galaxy evolution. High spatial resolution is clearly an advantage; \citet{Yuan2013} demonstrate how degraded resolution can dramatically bias gradient measurements. For typical-size galaxies at high redshift, $\lesssim1$ kpc resolution is essential to sample the scale radius and reliably measure gradient slopes. The use of single strong-line metallicity indicators such as [N~{\sc ii}]$/$H$\alpha$\ \citep[e.g.,][]{Pettini2004} is another potential problem. Multiple emission line diagnostics are necessary to confirm variations in metallicity \citep[e.g.,][]{Jones2013} and to distinguish H~{\sc ii}\ regions from AGN or shock excitation, which in many cases are non-negligible \citep{Wright2010,Yuan2012,Newman2014} and would bias the inferred gradients. In essence we require larger samples with high spatial resolution and a reliable suite of metallicity and excitation diagnostics in order to resolve the discrepancy in existing measurements. Here we present initial results from the {\it Grism Lensed-Amplified Survey from Space} (GLASS) in which we measure metallicities at $z\simeq2$ based on 5 nebular emission lines with resolution as fine as 200 pc. As with \citet{Jones2013} this is enabled by a combination of strong gravitational lensing, broad wavelength coverage, and diffraction-limited data. These results demonstrate the potential of the full GLASS survey to obtain dozens of such measurements at $1.3\lesssim z \lesssim2.3$ and thus reliably determine the average evolution of metallicity gradients over the past 11 Gyr. Throughout this paper we adopt a flat $\Lambda$CDM cosmology with $\Omega_{\Lambda}=0.7$, $\Omega_{\rm M}=0.3$, and H$_{\hbox{\scriptsize 0}}$$=70$ km~s$^{-1}$~Mpc$^{-1}$. At $z=1.85$, 3.5 Gyr has elapsed since the big bang and 1 arcsecond corresponds to 8.4 kpc (c.f. 3.5 Gyr and 8.7 kpc/arcsecond for the \citealt{Planck2013} cosmology). All stellar masses (M$_*$) and star formation rates (SFR) correspond to a \cite{Chabrier2003} initial mass function. Unless stated otherwise, "metallicity" refers to the gas-phase oxygen abundance. \section{A triple galaxy system at $z=1.855$} The first GLASS observations of MACS J0717+3745 (hereafter J0717) revealed three different strongly lensed sources at the same redshift $z=1.855\pm0.01$ (arc systems 3, 4, and 14 in \citealt{Schmidt2014}). The grism observations cover three magnified images of each galaxy (e.g., arcs 3.1, 3.2, and 3.3 are multiple images of the same source) and the details of each image are given in \cite{Schmidt2014}. The gravitational lensing model discussed in Section~\ref{sec:model} indicates projected galaxy separations of 50 kpc for pairs 4--14, 150 kpc for 3--14, and 200 kpc for 3--4. Although the grism redshifts cannot constrain their line-of-sight separation to better than a few Mpc, Keck/LRIS spectra of arcs 3.2 and 14.1 \citep{Limousin2012} and a Keck/MOSFIRE spectrum of arc 4.1 \citep{Schmidt2014} indicate consistent redshifts, corresponding to $\lesssim1$ Mpc in the Hubble flow. Therefore all three galaxies appear to be physically associated. We argue in Section~\ref{sec:results} that arcs 4 and 14 are likely affected by significant forces from gravitational interaction on the basis of their physical properties, however we caution that their true 3-D separation is uncertain. The fortuitous inclusion of these systems within {\it HST}'s field of view, and the availability of several strong emission lines in the grism spectra, provide a laboratory for studying how gas and heavy elements cycle within galaxies and their surrounding medium in an overdense environment. \section{Observations and data analysis}\label{sec:data} Details of the GLASS survey design and grism spectroscopy are given in \citet{Schmidt2014}. Briefly, J0717 was observed at two different position angles separated by approximately 100 degrees. The total exposure time is 10 orbits in the G102 grism and 4 orbits in the G141 grism. Data from each position angle are interlaced to produce 2D grism spectra and direct images with a scale of 0\farcs065 per pixel, Nyquist sampling the point spread function. Direct images are used to estimate continuum spectra by dispersing each pixel from individual objects detected in {\sc SExtractor} \citep{Bertin1996} segmentation maps; these are used to guide the extraction of individual object spectra and to subtract an estimate of the contamination from nearby objects. The critical analysis method for this work is extraction of emission line maps. For each object of interest, we use the direct image to construct a segmentation map and shift this to the expected wavelength of strong emission lines using spectroscopic redshifts from the grism data (Figure~\ref{fig:spec2d}). We extract two-dimensional maps of emission line flux and uncertainty for the region defined by the segmentation map, and correct for residual background structure by subtracting a corresponding map at a nearby line-free wavelength (e.g., a rest-frame 4750 \AA\ map is typically used to correct H$\beta$\ for imperfect continuum and contamination estimates). Data affected by strong contamination are masked and treated as missing. We align the emission line maps extracted from both position angles using bilinear interpolation, and combine them with an inverse-variance weighting. Example emission line maps are shown for arc 4.1 in Figure~\ref{fig:spec2d_arc4}. The H$\beta$\ and [O~{\sc iii}]\ emission lines require special treatment in cases where the spatial extent is sufficiently large that these lines overlap. In all cases there is a region of pure [O~{\sc iii}]\ $\lambda$5007 emission at the rightmost edge of the blend (of $\simeq 6$ pixels or $0\farcs4$ in the dispersion direction for $z=1.855$). We use the intrinsic flux ratio $f_{5007}/f_{4959} = 3$ to subtract the corresponding [O~{\sc iii}]\ $\lambda$4959 flux, which leaves another region of pure [O~{\sc iii}]\ $\lambda$5007 emission. We iterate this process to completely de-blend the [O~{\sc iii}]\ doublet, resulting in separate $\lambda$4959 and $\lambda$5007 emission line maps. Additionally we de-blend [O~{\sc iii}]\ $\lambda$4959 and H$\beta$, although we cannot completely de-blend the lines in cases where [O~{\sc iii}]\ $\lambda$5007 and H$\beta$\ overlap. For $z=1.855$ this occurs whenever the source size is $\geq 1\farcs4$ in the dispersion direction (for example, the spectrum of arc 14.1 shown in Figure~\ref{fig:spec2d}). In such cases we treat the region where [O~{\sc iii}]\ $\lambda$5007 and H$\beta$\ overlap as missing data. \begin{figure} \includegraphics[width=0.48\textwidth]{figure_arcs_spectra_v2.pdf} \caption{ \label{fig:spec2d} Grism spectra for the arcs of interest show multiple spatially extended emission lines which can be used to derive metallicity maps. From left to right, each row shows the F140W direct image, G102 grism spectrum, and G141 grism spectrum. In all cases the estimated contamination has been subtracted from the spectra. Strong source continuum is apparent for arc 14.1. Red contours show the object segmentation maps derived using the direct images, and mapped to emission lines of interest in the grism spectra using the redshift of each source (left to right: [O~{\sc ii}]\ $\lambda\lambda$3727, [Ne~{\sc iii}]\ $\lambda$3869, H$\gamma$, H$\beta$, [O~{\sc iii}]\ $\lambda$4959, [O~{\sc iii}]\ $\lambda$5007). The redshift is $z=1.855$ for all cases shown here. } \end{figure} \begin{figure} \includegraphics[width=0.5\textwidth]{figure_arc41_linemaps.pdf} \caption{ \label{fig:spec2d_arc4} {\it HST}\ image (top right; RGB: WFC3/F160W, WFC3/F110W, ACS/F814W) and emission line maps of arc 4.1. The typical flux uncertainty in each pixel is $2 \times 10^{-18}$ $\mathrm{ergs}~\mathrm{s}^{-1}~\mathrm{cm}^{-2}$ (1$\sigma$). } \end{figure} \section{Gravitational lensing model}\label{sec:model} An accurate gravitational lensing model is essential for reconstructing the source plane morphology of lensed galaxies, and for combining the information from the multiple images. As one of the Frontier Fields, the gravitational lensing potential of J0717\ has been modeled by several groups using a variety of techniques, and the results are publicly available\footnote{http://www.stsci.edu/hst/campaigns/frontier-fields/}. We compared the results of all models for which we could derive deflection angle maps at the redshift of interest (those of Brada{\v c}, CATS, Sharon, and Zitrin) to assess which is best suited to the purposes of this work. We note that those models are aimed at producing a global description of the cluster and therefore their accuracy in the vicinity of the lensed images of interest is expected to vary significantly between them. The Sharon version 2 model \citep{Johnson2014} produced the best results for our images, yielding the most precise inversion. The precision of the inversion was evaluated by comparing for each set of multiple images the promixity of the inferred source position, and the agreement beetween source plane flux and morphology. However, even the best global model produced significant residual differences between the reconstructed sources, requiring an additional step, as described in the next paragraph. In order to take full advantage of the multiple images of each arc system, we have developed a novel technique to align all images in the source plane. This allows us to reduce the lensing-related uncertainties and combine the data from multiple images to increase the signal-to-noise ratio of emission line maps. The full details of our methodology will be presented by Wang et al. (2014, in prep); here we give only a brief overview. Essentially, we are considering the global cluster model as an approximate first solution and we are seeking corrections to the potential to improve the reconstruction. We assume that the corrections are small and can thus be described by a local correction to the lensing potential up to the first two orders of derivatives. The first order term consists of a correction to the deflection angle for each image. The second order term yields corrections to the shear and convergence. The procedure has thus five free parameters per image. The optimal parameters are found by requiring the source plane reconstructions of each set of multiple images to be as similar to each other as possible. We note that a direct byproduct of this formalism is a correction to the magnification of each image. We have applied this method to the arc systems 3, 4, and 14, using the least distorted (i.e., least magnified) image of each system as a reference. Figure~\ref{fig:lenscorr} shows multiple images of arc 14 reconstructed with and without the lens model corrections to illustrate the advantage. The source plane positions are offset by $\simeq 0\farcs3 \simeq 3$ kpc using the original model, while the model corrections produce consistent positions and morphologies for all three reconstructed images. The second order correction is thus sufficient for our purposes. In all cases the correction is relatively small, amounting to $<5$\% of the total lensing deflection angle. In the following analysis we combine emission line maps from multiple images in the source plane in order to increase the measurement precision. However, including less highly magnified images provides a marginal improvement at the cost of degraded spatial resolution. We therefore use arcs 3.1$+$3.2, arc 4.1 only, and arcs 14.1$+$14.3 to optimize resolution and sensitivity. Final results are consistent with measurements from individual images, as well as with results derived from the original lens model. Combining multiple images improves the precision in metallicity gradients by an impressive factor of $\simeq2\times$ for arcs 3 and 14 by enabling finer spatial sampling and detection of more extended low surface brightness emission. We derive a conservative uncertainty of 13\% RMS in the magnification factors (prior to second order correction) by comparing measured flux ratios of multiple images with model-predicted magnification ratios. This error is not propagated in the following analysis, however it is small compared to other sources of uncertainty and has no effect on the results. \begin{figure} \includegraphics[width=0.5\textwidth]{14_durows_fin.pdf} \caption{ \label{fig:lenscorr} Demonstration of the lens model corrections. The panels in the top row are source plane reconstructions given by the original Sharon version 2 model, whereas those in the bottom row show the improved reconstructions after our correction is applied. Note that the original and corrected arc 14.2 is identical since it is used as a reference. In all panels, the gray scale represents the surface brightness contrast and the red contour shows an isophotal radius measured for the combined arc. Here we show the process for arc 14 as an illustration. The corrections for arcs 3 and 4 give similar results.} \end{figure} \begin{figure} \includegraphics[width=.3\columnwidth]{figure_arc3_source_fin.pdf} \hspace{0.03\columnwidth} \includegraphics[width=.3\columnwidth]{figure_arc4_source_fin.pdf} \hspace{0.03\columnwidth} \includegraphics[width=.3\columnwidth]{figure_arc14_source_fin.pdf} \caption{ \label{fig:sourceplane} Source plane morphologies of arcs 3, 4, and 14 (from left to right). Each panel shows contours of constant de-projected galactocentric radii at intervals of 1 kpc, derived in Section~\ref{sec:properties}. {\it Top row:} {\it HST}\ image (RGB: WFC3/F160W, WFC3/F110W, ACS/F814W). {\it Middle:} stellar mass maps derived from spatially resolved {\it HST}\ photometry. All galaxies have smooth, centrally peaked stellar mass profiles with no significant secondary peaks. {\it Bottom:} gas-phase metallicity maps. } \end{figure} \section{Physical properties}\label{sec:properties} \subsection{Stellar mass}\label{sec:mstar} We use the stellar population synthesis code FAST version 1.0 \citep{Kriek2009} to fit the resolved spectral energy density (SED) of each galaxy of interest. For $z=1.855$, the rest-frame UV through optical SEDs are well constrained by broad-band {\it HST}\ photometry taken as part of the CLASH survey \citep{Postman2012}. We utilize the ACS/F435W, ACS/F606W, ACS/F814W, and WFC3/F125W filters which sample nearly the full rest-frame wavelength range from 1250--4900 \AA. This set of filters is chosen to provide the widest possible wavelength coverage while avoiding contamination from the strong emission lines Ly$\alpha$, [O~{\sc ii}], and [O~{\sc iii}]\ which can significantly affect the broad-band photometry and derived stellar population properties. In particular, [O~{\sc iii}]\ emission accounts for $\sim$25\% of the total WFC3/F160W flux (an increase of 0.3 magnitudes) and [O~{\sc ii}]\ significantly affects the WFC3/F105W flux for the galaxies discussed here. Our methodology is as follows. We first align all images with the GLASS F105W direct image and smooth to a common point spread function of 0\farcs2 FWHM. The broad-band fluxes in each pixel are fit with a \citet{Bruzual2003} stellar population library, \cite{Chabrier2003} initial mass function, Milky Way dust attenuation law, stellar ages between 5 Myr and the age of the universe at the galaxy's redshift, and an exponentially declining star formation history with $\tau = 10^{7}-10^{10}$ yr. The quantity of greatest interest is the derived stellar mass surface density, which is the most robust parameter. Other stellar population parameters (SFR, age, extinction, etc.) are obtained simultaneously albeit with larger uncertainty. We have repeated the SED analysis using additional broad-band filters corrected for emission line contamination using the flux maps described in Section~\ref{sec:data}, verifying that this produces consistent results. Total stellar masses derived from integrated photometry and corrected for magnification are listed in Table~\ref{tab:arcs}. \subsection{Morphology}\label{sec:morphology} Morphological information is critical for measuring accurate metallicity gradients, as the radial coordinate depends on a galaxy's central position and inclination. Ideally the dynamical center, major axis orientation, and inclination would be constrained from kinematics as has been done for previous work \citep[e.g.,][]{Jones2013}, but we lack kinematic data. Instead we derive estimates of these quantities by assuming that the stellar mass surface density derived in Section~\ref{sec:mstar}, $\Sigma_*$, is elliptically symmetric. We reconstruct $\Sigma_*$ using the lens model (Section~\ref{sec:model}) and fit for the centroid, orientation, and axis ratio. Only the most highly magnified image in each arc system is used in order to maximize the spatial resolution. In the following analysis we adopt the galaxy center and inclination such that contours of $\Sigma_*$ trace contours of constant de-projected radius. Source plane $\Sigma_*$ distributions and best-fit ellipses are shown in Figure~\ref{fig:sourceplane}. All galaxies exhibit smooth, centrally peaked stellar mass profiles with contours that are fit by ellipsoids. Within the stellar mass uncertainty, we find no evidence for ongoing late-stage major mergers which could manifest as a secondary peak in the stellar mass density. \subsection{Metallicity and nebular extinction}\label{sec:z} We use the strong line ratio calibrations presented by \cite{Maiolino2008} to estimate gas-phase oxygen abundance (expressed as $12+\log{\mathrm{O/H}}$) and nebular extinction A(V) from measured [O~{\sc ii}], [Ne~{\sc iii}], H$\gamma$, H$\beta$, and [O~{\sc iii}]\ fluxes. An advantage of these lines is that they directly trace the oxygen abundance, as opposed to [N~{\sc ii}]-based diagnostics which may suffer from redshift-dependent systematic errors arising from evolution in the N/O ratio or other effects \citep[e.g.,][]{Shapley2014,Steidel2014}. We use a \cite{Cardelli1989} extinction curve with R$_{\rm V} = 3.1$, noting that the choice of R$_{\rm V}$ has no significant effect on the derived metallicity (typically $<0.1$ dex for $2\leq {\rm R_V} \leq 5$). To further constrain A(V), we impose $\frac{\rm{H}\gamma}{\rm{H}\beta} = 0.47$ as expected for Case B recombination and typical H~{\sc ii}\ region conditions \citep[e.g.,][]{Hummer1987}. Metallicity and extinction are derived from a $\chi^2$ statistic constructed from all available diagnostics: $$\chi^2 = \sum_R \left[ \frac{\Delta R}{\sigma(R)} \right]^2.$$ Here $\Delta R$ is the difference between predicted and observed de-reddened flux ratios at a given A(V) and $12+\log{\mathrm{O/H}}$, and the uncertainty $\sigma(R)$ includes RMS scatter in each calibration added in quadrature with measurement uncertainty. The diagnostic ratios used are R$_{23}$ $=$ ([O~{\sc ii}]+[O~{\sc iii}])/Hb, [O~{\sc iii}]/[O~{\sc ii}], [O~{\sc iii}]/Hb, [O~{\sc ii}]/Hb, [Ne~{\sc iii}]/[O~{\sc ii}], and H$\gamma$/H$\beta$. We adopt the best fit as the values that minimize $\chi^2$, and 1$\sigma$ uncertainty from the extrema at which $\chi^2$ differs by 1 from the minimum value. This is slightly smaller than the formal uncertainty because of covariance between various metallicity diagnostics; we have checked that a more careful treatment does not affect the conclusions. We limit the solutions to A(V)~$=0-3$ and $12+\log{\mathrm{O/H}}$~$=7.0-9.3$ \citep[equivalent to $0.02-4.1\times$ the solar abundance of $12+\log{\mathrm{O/H}}$~$=8.69$;][]{Allende2001}. While the formal best-fit solutions are occasionally beyond this range, all such cases are poorly constrained and we view such extreme conditions as highly unlikely. Although the extinction is poorly constrained for most individual pixels, SED fitting (Section~\ref{sec:mstar}) and averaged H$\beta$$/$H$\gamma$\ ratios indicate relatively low extinction toward both the stars and H~{\sc ii}\ regions, with best-fit A(V)$=0-0.6$ and permitted A(V)$<1.1$ (1$\sigma$) in the dustiest case. Fortunately, uncertainty in extinction has little effect on the derived metallicity since many diagnostic line ratios are relatively insensitive to reddening. Maps of the source plane gas-phase metallicity are shown in Figure~\ref{fig:sourceplane}. The emission line maps are binned to spatial scales of $\simeq300-500$ pc and the metallicity is fit for all such pixels where at least one emission line is detected at $\geq 5\sigma$ significance. All sources are well resolved with measurements extending over $\simeq3$ resolution elements for arc 3, and $\geq 6$ resolution elements for the others. The metallicity is typically sub-solar with most pixels having best-fit $12+\log{\mathrm{O/H}}$\ in the range 7.5$-$8.5, corresponding to flux ratios $\frac{\rm [O~\textsc{iii}]}{\rm{H} \beta} > 2.5$ and $\frac{\rm [O~\textsc{iii}]}{\rm [O~\textsc{ii}]} > 1$. Figure~\ref{fig:sourceplane} also shows several pixels for which the best-fit metallicity is at the extreme ends of the permitted range; these usually correspond to noise artifacts and/or single-line detections and are therefore unreliable. In addition to the spatially resolved analysis, we calculate galaxy-averaged metallicity and extinction by applying the same analysis to total emission line fluxes measured from integrated 1-D spectra. The results agree with the average of individual pixel values (Table~\ref{tab:arcs}), providing a valuable sanity check. The slightly lower average metallicity of individual pixels c.f. integrated flux is likely due to the requirement of a 5$\sigma$ detection of [O~{\sc iii}]\ (the brightest emission line), which induces a bias toward lower metallicity. Extinction-corrected Balmer line fluxes derived from this analysis are used to determine star formation rates in Section~\ref{sec:sfr}. \subsection{Metallicity gradient}\label{sec:gradients} Gas-phase metallicity gradients are derived from the morphology and metallicity analyses described in Sections~\ref{sec:morphology} and \ref{sec:z}, respectively. In essence we measure the mean metallicity as a function of galactocentric radius in the source plane, utilizing knowledge of the major axis and inclination of each galaxy. We consider two approaches based on (1) individual pixels and (2) radial binning. For the first method, we compute the gradient $\frac{\Delta Z}{\Delta R}$ from a linear fit to the metallicity vs. de-projected radius of individual pixels (those shown in Figure~\ref{fig:sourceplane}): \begin{equation}\label{eq:gradient} $$12+\log{\mathrm{O/H}}$$\, = Z_0 + \frac{\Delta Z}{\Delta R} R. \end{equation} This functional form is a good fit to the data and we are unable to distinguish more complex behavior given the measurement uncertainties. Pixels for which the allowed (1$\sigma$) range of $12+\log{\mathrm{O/H}}$\ extends to $\leq7$ or $\geq9.3$ are excluded from the fit as these are generally unreliable and have large uncertainty ($\gtrsim 1$ dex), although including these data gives consistent results. For the second method we bin the emission line flux in annular apertures for each source, calculate metallicity from the total flux in each aperture, and compute the gradient using Equation~\ref{eq:gradient}. The range of radii for each annulus is chosen to provide a $\simeq10\sigma$ detection of [O~{\sc iii}], which is uniformly the strongest emission line. We show the data and best-fit gradients in Figure~\ref{fig:gradients} and list the results in Table~\ref{tab:arcs}. Uncertainty in inclination and magnification are not included in these error estimates, but these effects are small ($\lesssim 30$\% combined) compared to uncertainty in the metallicities. Both methods of calculating the metallicity gradient use the same diagnostics applied to the same data set and give consistent results. The pixelated approach is similar to local galaxy measurements which are based on individual H~{\sc ii}\ regions \citep[e.g.,][]{Vila-Costas1992}, and the radial binning method is equivalent to most measurements reported for high redshift galaxies \citep{Yuan2011,Swinbank2012,Queyrel2012,Stott2014}. The primary differences are that the annular binning method is higher signal to noise ratio and is spatially complete, whereas low surface brightness regions are either noisy or excluded from the pixelated method. We therefore consider the annular binning method to be more robust. \begin{figure*} \includegraphics[width=.33\textwidth]{figure_arc3_gradient_bin_fin.pdf} \includegraphics[width=.33\textwidth]{figure_arc4_gradient_bin_fin.pdf} \includegraphics[width=.33\textwidth]{figure_arc14_gradient_bin_fin.pdf} \caption{ \label{fig:gradients} Radial metallicity gradients in each galaxy. Grey points are measured from individual source plane pixels shown in Figure~\ref{fig:sourceplane}, and thick black points show the results for flux summed within radial annuli. Linear gradient fits to the individual pixels and radial bins are shown as dashed and solid lines, respectively. Both methods give consistent results.} \end{figure*} \subsection{Star formation rate}\label{sec:sfr} Star formation rates are derived from Balmer emission using standard methods. We convert the extinction-corrected total H$\beta$\ flux (Section~\ref{sec:z}) to SFR following \cite{Kennicutt1998}, divide by 1.7 to convert to a \cite{Chabrier2003} IMF, and divide by the lensing magnification to recover the intrinsic SFR of each galaxy. Results are given in Table~\ref{tab:arcs} including uncertainty from measurement noise and extinction. Additional uncertainties associated with the choice of $R_V$ and underlying stellar absorption (for which we make no correction) are negligible. \subsection{AGN contamination} In calculating metallicities and SFR we have assumed that all nebular emission arises in H~{\sc ii}\ regions. The presence of shocks or AGN can bias these results even for very weak nuclear sources and we therefore check for any possible contamination. The global flux ratios of each galaxy, as well as the individual annular bins shown in Figure~\ref{fig:gradients}, are consistent with H~{\sc ii}\ regions. However in all cases the line ratios fall in the region occupied by both H~{\sc ii}\ and AGN excitation \citep[for example in the "blue" diagnostic diagram of [O~{\sc iii}]/Hb vs. [O~{\sc ii}]/Hb, e.g.,][]{Lamareille2010}, and we lack a low-excitation line such as [N~{\sc ii}]\ or [S~{\sc ii}]\ which would permit more reliable classification \citep[e.g.,][]{Jones2013}. Publicly available Chandra/ACIS data\footnote{Available at http://cxc.cfa.harvard.edu/cda/} show no X-ray detection for any of the arcs, and all three arcs have moderate-resolution Keck spectra which show no sign of AGN features \citep{Schmidt2014,Limousin2012}. We note that the central resolution element of arc 14 shows marginal evidence ($\sim2\sigma$) of an elevated [O~{\sc iii}]$/$H$\beta$\ ratio as expected for AGN \citep[e.g.,][]{Wright2010,Newman2014}. \cite{Trump2011} have associated this signal with a significant AGN fraction in stacks of galaxies of similar stellar mass and redshift whose individual X-ray luminosities are below the detection threshold, although it is not a reliable indicator of AGN in individual galaxies. To summarize, we find no conclusive evidence of AGN but cannot rule out a possible low-level contribution to the emission line flux from a nuclear source. The expected effect of such activity would be to lower the derived central metallicity, with no significant effect on SFR or stellar masses. \begin{deluxetable*}{lcccccccc} \tablecolumns{8} \tablewidth{0pt} \tablecaption{Source properties\label{tab:arcs}} \tablehead{ \colhead{ID} & \colhead{$z$} & \colhead{log M$_*$} & \colhead{SFR} & \colhead{$12+\log{\mathrm{O/H}}$\tablenotemark{1}} & \colhead{$12+\log{\mathrm{O/H}}$\tablenotemark{2}} & \colhead{$\Delta \rm{Z} / \Delta \rm{R}$\tablenotemark{3}} & \colhead{$\Delta \rm{Z} / \Delta \rm{R}$\tablenotemark{4}} \\ \colhead{} & \colhead{} & \colhead{[log $\mathrm{M}_{\sun}$]} & \colhead{[$\mathrm{M}_{\sun}~\mathrm{yr}^{-1}$]} & \colhead{} & \colhead{} & \colhead{[dex kpc$^{-1}$]} & \colhead{[dex kpc$^{-1}$]} \\ } \medskip \startdata arc 3 & 1.855 & 7.2$\pm$0.4 & 1.8$^{+1.0}_{-0.8}$ & 7.56$\pm$0.37 & 7.75$^{+0.25}_{-0.19}$ & -0.07$\pm$1.06 & -0.29$\pm$0.25 \\ arc 4 & 1.855 & 9.1$^{+0.2}_{-0.1}$ & 29$^{+21}_{-14}$ & 8.20$\pm$0.34 & 8.35$^{+0.11}_{-0.13}$ & -0.03$\pm$0.03 & -0.05$\pm$0.05 \\ arc 14 & 1.855 & 8.8$^{+0.2}_{-0.1}$ & 3.3$^{+3.2}_{-0.9}$ & 8.11$\pm$0.31 & 8.30$^{+0.15}_{-0.23}$ & 0.34$\pm$0.41 & 0.28$\pm$0.29 \\ \enddata \tablenotetext{1}{Average and RMS scatter of individual pixels} \tablenotetext{2}{Best fit and 1$\sigma$ uncertainty derived from integrated spectrum} \tablenotetext{3}{Derived from individual pixels} \tablenotetext{4}{Derived from total flux in radial apertures} \end{deluxetable*} \section{Results and discussion}\label{sec:results} \subsection{Global properties} In this section we examine the GLASS sample studied here in the context of the general $z\simeq1.8$ galaxy population, as a prelude to discussing the implications of our results for galaxy evolution. Figure~\ref{fig:properties} summarizes the demographic properties compared to larger published samples at similar redshift. Comparison samples are carefully chosen to minimize systematic effects: the narrow range of redshift mitigates any evolutionary trends, all data points are calculated with the same IMF and stellar population templates, and all stellar masses account for the contribution of emission lines to broad-band photometry. Ignoring this last effect would result in higher stellar masses by $\simeq0.15$ dex for arcs 4 and 14 and $\simeq0.6$ dex for arc 3, in excellent agreement with the mass-dependent estimates by \cite{Whitaker2014}. In lieu of metallicity, we use [O~{\sc iii}]$/$[O~{\sc ii}]\ flux ratios for a clear comparison of observable properties. While this ratio is sensitive to metallicity, ionization parameter, and reddening, all of these quantities are intrinsically correlated and vary monotonically with [O~{\sc iii}]$/$[O~{\sc ii}]\ \citep[e.g.,][]{Maiolino2008,Dopita2013}. Furthermore, this ratio offers the lowest statistical uncertainty and the greatest sensitivity to metallicity of the available strong-line diagnostics in the sub-solar regime. We can therefore directly compare {\em relative} oxygen abundances between different samples, noting that secondary parameter dependences result in $\simeq0.25$ dex RMS scatter between metallicity and [O~{\sc iii}]$/$[O~{\sc ii}]. In comparison to the overall population, the galaxies studied here have SFRs which are a factor of $\geq3\times$ above the "main sequence" locus of SFR and M$_*$ (Figure~\ref{fig:properties}). Such high relative SFRs are often induced by gravitational interactions, and indeed \cite{Stott2013} show that $\sim50$\% of sources at this SFR, M$_*$, and $z$ are merger events. The metallicities inferred from [O~{\sc iii}]$/$[O~{\sc ii}]\ are consistent with comparison samples of similar SFR and M$_*$, as are the nebular extinction values. The GLASS data are also consistent with a steep mass-metallicity relation as found by \cite{Henry2013}, although the scatter in Figure~\ref{fig:properties} is large. To summarize, the global properties are typical of low-mass starbursts at $z\simeq2$ which are frequently associated with galaxy interactions. \begin{figure} \includegraphics[width=\columnwidth]{figure_ms_fin.pdf} \\ \includegraphics[width=\columnwidth]{figure_o32_fin.pdf} \caption{ \label{fig:properties} Global demographic properties of the GLASS arcs. For comparison we also show average values of larger samples at similar redshift from WISP \citep[][H13]{Henry2013} and 3D-HST \citep[][W14]{Whitaker2014} data as well as individual lensed galaxies reported by \citet[][B13]{Belli2013}. {\it Top:} SFR versus stellar mass. The arcs studied here have significantly higher specific SFR compared to the "main sequence" at $z\simeq1.8$ (shown by the W14 data), consistent with expectations for excess star formation triggered by gravitational interactions among these systems. {\it Bottom:} comparison of emission line ratios versus stellar mass. Open symbols show the effect of dust attenuation measured by \cite{Henry2013}; all other data are uncorrected for reddening. The arcs studied here have typical [O~{\sc iii}]$/$[O~{\sc ii}]\ ratios for their redshift, stellar mass, and SFR. The right axis shows equivalent metallicity corresponding to [O~{\sc iii}]$/$[O~{\sc ii}]\ for the \cite{Maiolino2008} calibration used in this work. The GLASS data are consistent with a steep mass-metallicity relation as found by \cite{Henry2013}, and notably extend to an order of magnitude lower in stellar mass compared to previous studies at $z\simeq2$.} \end{figure} \subsection{Metallicity gradient evolution} We now turn to a discussion of the metallicity gradient of arc 4 and its evolution with redshift. Figure~\ref{fig:evolution} compares the GLASS data with similar measurements as a function of redshift. Arcs 3 and 14 are not included because their gradients are poorly constrained, due to their small physical sizes and lower signal to noise ratio. In all three cases the gradients are consistent with zero at the $1\sigma$ level. We restrict the comparison sample to measurements with $\lesssim1$ kpc resolution, noting that even kpc resolution results in artificially flat inferred gradients for typical high redshift galaxies \citep[e.g.,][]{Yuan2013,Stott2014}. We show original published values for all comparison data with the caveat that they are derived using different strong-line calibrations; this has minimal effect on our conclusions since metallicity gradients are relatively insensitive to the choice of calibration \citep{Jones2013}. The lensed $z=0.8$ source described by \cite{Frye2012} is also included with a previously unpublished gradient slope of $0.01\pm0.03$ dex kpc$^{-1}$ (B. Frye, private communication). However none of the comparison data include scatter in metallicity in the formal uncertainty, hence the error bar on arc 4 is the most conservative of those shown in Figure~\ref{fig:evolution}. Ideally we would construct a sample which corresponds to approximately the same galaxy population \citep[as was done in][]{Jones2013}, but we are presently limited by the available data at high redshift. The result is that arc 4 is expected to have $\sim0.4$ dex lower stellar masses at a given redshift than average sources in Figure~\ref{fig:evolution}, which are approximately Milky Way analogs \citep[based on abundance matching; e.g.,][]{Behroozi2013}. Nonetheless this provides a useful comparison. Figure~\ref{fig:evolution} demonstrates that interacting galaxies (shown as open symbols) have flatter gradients than isolated counterparts. This is most evident in the lensed and $z=0$ samples which have the best spatial resolution. Here we include arc 4 as an interacting system on the basis of its proximity to arc 14 and its enhanced SFR. Measurements of nearby galaxies have shown dramatically flattened gradients in pairs with the same projected separation as arcs 4--14 \citep[][and references therein]{Rich2012} which further motivates this classification. The relatively shallow metallicity gradient of arc 4 is therefore fully consistent with expected gravitational interaction. The gradient of arc 14 is also consistent with a near-zero value expected in this scenario. Arc 3 may have a shallow gradient as well, although this is not necessarily expected given its larger projected distance. An alternative cause for shallow gradients measured in the GLASS data is metal mixing by strong feedback. To illustrate this, Figure~\ref{fig:evolution} shows evolutionary tracks for two simulations with different sub-grid feedback prescriptions of an otherwise identical Milky Way-like galaxy discussed by \cite{Gibson2013}, and we now briefly summarize their findings. The "normal" feedback simulation \citep[MUGS;][]{Stinson2010} results in relatively steep metallicity gradients of 0.2--0.3 dex\,kpc$^{-1}$ at $z\simeq2$ which flatten at later times. The MaGICC simulation \citep{Brook2011} employs an enhanced feedback scheme, in which galactic outflows are more effective at removing metal-enriched ISM material which is re-accreted preferentially in outer regions. This mixing of the ISM results in shallower gradients of $<0.05$ dex\,kpc$^{-1}$ which become marginally steeper with time. Other groups have found essentially identical behavior with various feedback mechanisms which act to redistribute energy and ISM material throughout galactic disks \citep[e.g.,][]{Yang2012,Angles-Alcazar2014}. The enhanced feedback results in Figure~\ref{fig:evolution} are in excellent agreement with the shallow gradient measured for arc 4 as well as for arc 14. Arc 3 is expected to have a steeper negative gradient than shown by these models due to its lower mass \citep[e.g.,][]{Few2012}, but we are unable to distinguish between a shallow and steep negative gradient within the uncertainty. The overall metallicity can in principle help to identify the cause of shallow gradients: rapidly recycled outflows should produce a signature of higher metallicity, whereas gravitational interactions should have an opposite effect \citep[e.g.,][]{Rich2012}. We note that \cite{Kulas2013} report evidence for rapid recycling in a protocluster at $z=2.3$ based on stacked spectra, and we might expect a similar observational signature in the overdense region studied here. However, the expected differences are small \citep[$\sim0.1$ dex;][]{Torrey2012,Kulas2013} compared to measurement uncertainties so we are unable to distinguish these scenarios without additional complementary data. In any case the shallow gradient in arc 4 suggests that metals are efficiently mixed throughout the ISM on scales of several kpc. \begin{figure} \includegraphics[width=0.45\textwidth]{figure_gradient_evolution_fin.pdf} \caption{ \label{fig:evolution} Evolution of metallicity gradients with redshift. We compare the gradient of arc 4 from this work with published measurements at high redshift including other lensed galaxies (\citealt{Jones2013}, J13; \citealt{Yuan2011}, Y11; \citealt{Frye2012}, F12), non-lensed galaxies observed with adaptive optics \citep[][S12]{Swinbank2012}, an average of local gradients reported by \citet[][R10]{Rupke2010b}, and the Milky Way's metallicity gradient evolution measured from planetary nebulae \citep[][M03]{Maciel2003}. Solid and hollow symbols denote isolated disks and interacting (or merging) galaxies, respectively, to show that interacting galaxies have flatter gradients (closer to zero) on average in the lensed and $z=0$ samples. We additionally show results of two different feedback schemes in otherwise identical simulations, described in \cite[][G13; the galaxy shown is g15784]{Gibson2013}. The standard feedback results are similar to the Milky Way and isolated lensed galaxies, while enhanced feedback leads to shallower gradients at high redshifts and is good agreement with the measurement of arc 4 in this work. } \end{figure} \subsection{A starbursting dwarf galaxy at $z\simeq2$} One of the most striking aspects of Figure~\ref{fig:properties} is the low stellar mass of arc 3, M$_* = 1.5 \times 10^7 \mathrm{M}_{\sun}$. This is an order of magnitude below previous metallicity studies at $z\simeq2$ and extends into the regime of Milky Way dwarf satellites. Since such galaxies have rarely been studied directly at high redshift \citep{Christensen2012,Stark2014,Atek2014}, we briefly consider the properties of arc 3 in the context of recent cosmological simulations and local group analogs. Both simulations and observations of nearby dwarf galaxies suggest that in the mass regime of arc 3, $\sim$50\% of its present-day stellar mass has already formed by $z=1.85$ \citep{Shen2014,Weisz2011}. In terms of mass it is therefore roughly analogous to the "Doc" simulation of \cite{Shen2014}, and to the Milky Way dwarf spheroidal Fornax \citep[e.g.,][]{Coleman2008}. This is supported by the overall oxygen abundance $12+\log{\mathrm{O/H}}$\,$=7.75$ (equivalent to [O/H]~$=-0.94$) which is in excellent agreement with the stellar $\alpha$-element abundance distribution of Fornax \citep{Kirby2011}. Figure~\ref{fig:gradients} indicates a gas-phase metallicity gradient of $-0.29\pm0.25$ which is also compatible with Fornax's {\em present-day stellar} metallicity gradient \citep[$-0.50\pm0.10$ dex\,kpc$^{-1}$;][]{Hendricks2014}. However the uncertainty is large, and these two quantities need not be identical if there is significant radial growth or dynamical evolution of stellar orbits. Such a steep gradient would imply that any dynamical interaction in arc 3 is insufficient for strong radial mixing, and this is consistent with \cite{Rich2012} given that the nearest detected neighbor (arc 14) is separated by 150 kpc in projection. We measure an instantaneous specific star formation rate of $\log{\rm SSFR (yr^{-1})} = -7.0 \pm 0.5$ for arc 3, i.e., a mass doubling time of only $10^{7\pm0.5}$ yr. However we caution that this is calculated assuming constant SFR over the previous $\sim20$ Myr, which is likely not the case: such a high value, as well as the high ratio of Balmer line flux to UV luminosity and the stellar population modeling described in Section~\ref{sec:mstar}, suggests a short duty cycle for the current burst of star formation. Such intense short starbursts are predicted by several simulations of dwarf galaxies \citep[e.g.,][]{Governato2012,Zolotov2012,Brooks2014,Shen2014}, and may be triggered by interactions or by gradual accumulation of a large gas reservoir. Bursty star formation histories are of particular interest in light of observations that the central regions of local group dwarf galaxies are less dense than predicted for $\Lambda$CDM cosmology in the absence of baryonic feedback \citep[the "too big to fail" problem;][]{Boylan-Kolchin2011}. These simulations have found that repeated bursts of star formation and strong feedback in dwarf galaxies can affect the gravitational potential, transforming dark matter density profiles from cusps to cores, and have been postulated to reconcile $\Lambda$CDM with observations of local group substructure \citep[although see also][]{Garrison-Kimmel2013}. The high SFR surface density of arc 3 is sufficient to drive outflows with high mass loading factors \citep{Newman2012} as required by these simulations. However we do not directly constrain the outflow properties, and direct measurements of outflowing gas are extremely challenging for the luminosity of arc 3. One means of making progress is to measure the mass-metallicity relation slope which is sensitive to the properties of stellar feedback. The data shown in Figure~\ref{fig:properties} are consistent with feedback in the form of strong energy-driven galactic winds \citep{Henry2013}, but of course a larger sample will be required to better constrain the mass-metallicity relation at M$_*\lesssim10^8 \mathrm{M}_{\sun}$. We expect to find several dozen such low-mass sources in the full GLASS survey, which will represent the first statistical constraint of chemical enrichment and feedback at high redshift in the mass regime where galaxies are in tension with $\Lambda$CDM. The present data for arc 3 are generally consistent with simulations in which bursty star formation and strong feedback significantly lower the central density of dwarf galaxies. \section{Conclusions}\label{sec:conclusions} This work presents a case study of galaxy evolution in a dense environment at $z=1.85$ from the GLASS survey, utilizing the spatial resolution of {\it HST}\ aided by gravitational lensing. We study three galaxies in close physical proximity (50--200 kpc projected separations) and measure their spatially resolved metallicities from {\it HST}\ grism spectra. Our main results are as follows: \begin{itemize} \item All three galaxies appear to be affected by interaction-induced gas flows, evidenced by their close proximity and specific star formation rates which are $\geq3\times$ above the median value for their M$_*$ and redshift. \item We measure a precise metallicity gradient for one galaxy and find a shallow slope compared to other sources at similar redshift, as expected from gravitational interactions. The gradients are also consistent with strong stellar feedback resulting in higher metallicity of accreted (recycled) gas, although we find no evidence of enhanced metallicity which should result from this effect. \item The stellar mass range extends an order of magnitude below previous metallicity surveys at $z\simeq2$. The data are consistent with a steep mass-metallicity relation found in previous work, and the full GLASS survey should yield a sufficient sample to measure metallicity trends in the range M$_* \simeq10^7-10^8 \mathrm{M}_{\sun}$. \item We present the first spatially resolved spectrophotometric analysis of a bona-fide dwarf galaxy at $z\simeq2$ with stellar mass M$_*=10^{7.2\pm0.4}\, \mathrm{M}_{\sun}$. It is analogous to Fornax in terms of its stellar mass, metallicity, and metallicity gradient. Its extremely high SSFR indicates a violent starburst, likely with a low duty cycle, consistent with the hypothesis that successive starbursts with strong feedback are responsible for flattening the central density profiles of dwarf galaxies. \end{itemize} Measuring accurate metallicity gradients at high redshift is a significant observational challenge which, when overcome, is expected to elucidate various aspects of the galaxy formation process. Observing lensed galaxies at $\lesssim0\farcs2$ resolution is the best way to obtain the requisite spatial sampling \citep{Yuan2013}. Previously only 5 such gradient measurements have been reported in the literature \citep{Jones2010,Jones2013,Yuan2011}, although we note that improvements to the Keck/OSIRIS spectrograph and adaptive optics system have enabled recent observations of a significantly larger sample (Leethochawalit et al. in prep). Our analysis of {\it HST}\ data provides further support for a scenario in which gravitational interactions and mergers cause a temporary flattening of gradients, while isolated galaxies at high redshift have steep negative metallicity gradients consistent with standard feedback models. More importantly, our results constitute a proof-of-concept that {\it HST}\ grism spectra obtained with the GLASS survey yield precise gradient slope measurements with conservative uncertainty $\simeq0.05$ dex\,kpc$^{-1}$ in good cases. This is enabled by the combination of broad wavelength coverage and high spatial resolution aided by gravitational lensing. We note that similar previous work has shown the benefits of this approach, including several studies of lensed galaxies observed with {\it HST}\ grisms \citep{Brammer2012, Frye2012, Whitaker2014b}. This paper presents the complete methodology developed for such measurements, including a significant improvement over previous lensing studies by combining data from multiple strongly lensed images (Section~\ref{sec:model}; Wang et al. in prep). From visual inspection of the current GLASS data, we identify $\simeq5-20$ galaxies per cluster at $z=1-3$ with multiple bright emission lines suitable for resolved metallicity analysis. We expect the full survey to yield high-quality gradient measurements (comparable to arc 4) for approximately 20 of these galaxies. (This metallicity gradient sample will most likely be composed of relatively massive galaxies with M$_* \gtrsim 10^9 \, \mathrm{M}_{\sun}$, while at the same time we expect to characterize the properties of a larger sample of low-mass galaxies with modest spatial information). We anticipate that such a large sample will conclusively establish whether metallicity gradient slopes are steeper or shallower at $z\simeq2$ compared to local descendants, and more generally how metallicity gradients evolve over time. This will have significant implications for sub-grid feedback prescriptions used in cosmological simulations \citep[e.g.,][]{Gibson2013,Angles-Alcazar2014}, which in turn inform our understanding of baryon cycling and chemical enrichment of the circum- and inter-galactic media, and the role of this baryon cycle in galaxy formation. \acknowledgements We thank R. Maiolino for providing the metallicity calibrations used in this work, and B. Gibson for providing metallicity gradient evolution tracks from simulations. We thank the anonymous referee for a constructive report which improved the clarity of this paper. T. Johnson and K. Sharon are acknowledged for providing the map of the phase angle. TAJ acknowledges support from the Southern California Center for Galaxy Evolution through a CGE Fellowship. We acknowledge support from NASA through grant HST-13459. TT acknowledges support by the Packard Foundation in the form of a Packard Research Fellowship and thanks the American Academy in Rome and the Observatory of Monteporzio Catone for their generous hospitality. This paper is based on observations made with the NASA/ESA Hubble Space Telescope, and utilizes gravitational lensing models produced by PIs Brada{\v c}, Ebeling, Merten \& Zitrin, Sharon, and Williams funded as part of the HST Frontier Fields program conducted by STScI. STScI is operated by the Association of Universities for Research in Astronomy, Inc. under NASA contract NAS 5-26555. The lens models were obtained from the Mikulski Archive for Space Telescopes (MAST).
\section{Introduction} We have seen a number of recent advancements to the theory of rank aggregation. This problem has a number of applications ranging from marketing and advertisements to competitions and election. The main question of rank aggregation is how to consistently combine various individual preferences. This type of data is frequently available to us: what webpage did a user select, who won the chess match, which movie did a user watch, etc.... All of these examples yield comparisons without explicitly revealing an underlying score. That is, only the preference is observed, not necessarily the strength of the preference (in the case of sports one might argue that the score indicates such a magnitude difference). Additionally, numeric scores have been shown to be inconsistent and subject to variations in calibration in various contexts. Given how natural the problem of rank aggregation is, there has been a wide body recent~\cite{Duchi10,ammarandshah} and classical work~\cite{Arrow,Bradley,Condorcet,Luce} to understand how to consistently combine preferences. However, all of these methods have a major drawback: they aim to find \emph{one} ranking. In many settings, various individuals will have separate preferences, and we wish to model those distinctions. For example, we might wish to provide personalized ads, search results, or movie recommendations on a per user basis. In standard contexts we assume that there is one consistent ranking that does well to approximate the behavior of all users, but these aggregation methods cannot model the discrepancies across users. Our goal is to understand how to analyze a method that has the flexibility to account for user differences and can be adaptive; that is, if there are no differences, then the method should have stronger performance guarantees. This task can be seen as rank aggregation analog to the standard collaborative filtering problem. \comment{ The matrix completion problem has received extensive extension over the past decade and recently a number of theoretical advances have been made in understanding the matrix completion problem. Recent work has aimed to generalize the types of observation models that we can considering in the matrix completion framework. One popular framework has been the so-called ``one-bit'' matrix completion problem~\cite{onebit} where we only observe a $+1$ or $-1$ at each entry. This model is also the effective one considered by Srebro et. al.~\cite{srebro} in much of their early work on matrix completion. Matrix completion can also be seen as an approach to the collaborative filtering problem~\cite{collabfiltering}.} While there have been significant theoretical advances in the understanding of collaborative filtering, or more generally matrix completion~\cite{CanRecCompletion,TsyCompletion,NegWaiCompletion}, there has been far less work in understanding how to perform the proposed type of collaborative ranking. Recent work has demonstrated that taking rankings into consideration can significantly improve upon rating prediction accuracy~\cite{eigenrank,param17,cofirank,Yi13}, thus it is a natural question to understand how such collaborative ranking methods might behave. One reason for this discrepancy is this theoretical understanding of single user rank aggregation is already a very challenging problem as discussed above. Whereas, single rating aggregation is trivial: take an average. Another, possibly more interesting distinction is in the amount of apparent information made available. In the standard matrix completion setting we have direct (albeit noisy) access to the true underlying ratings. Therefore, if the noise is sufficiently small, we could order the information into a list. On the other hand, in the collaborative ranking problem we never have direct access to the true signal itself and only observe relative differences. In some sense, this is a harder problem~\cite{shahetal14} owing to the fact that the comparisons are in themselves functions of the underlying ratings. When we are given, for example, $\personex$ ratings, then we can convert that to $\binom{\personex}{2}$ pairwise comparisons. This crude analysis seems to indicate that we would require far greater pairwise comparisons in order to recover the true underlying matrix. We will show that this increase in the number of examples is not required. In the sequel, we will show that under a natural choice model for collaborative ranking, the total number of comparisons needed to estimate the parameters is on the same order as the total number of explicit ratings observations required in the standard matrix completion literature. Thus, we demonstrate that collaborative ranking based pair-wise comparisons from a simple and natural model can yield very similar results as in the standard matrix completion setting. \paragraph{Past Work} As alluded to above there has been some work in understanding collaborative rankings and learning user preferences. The nuclear norm approach is fundamentally a regularized $M$-estimator~\cite{Neg09}. The application of the nuclear norm approach to collaborative ranking was first proposed by Yi et al.~\cite{Yi13}. There work showed very good empirical evidence for using such a nuclear norm regularized based approach. However, that work left open the question of theoretical guarantees. Other results also assume that the underlying ratings are in fact available. However, rather than inferring unknown ratings their goal is to infer unknown ranked preferences \emph{from} known ratings. That is, they wish to deduce if a user will prefer one item over another rather than guess what their ratings of that item might be~\cite{eigenrank,cofirank,bayesmatrix}. The work by by Weimer et. al.~\cite{cofirank} also uses a nuclear norm regularization, but that work assumes access to the true underlying ratings, while we assume access only to pairwise preferences. Other algorithms aggregate users’ ratings by exploiting the similarity of users by nearest neighbor search~\cite{empirical,param17}, low-rank matrix factorization ~\cite{matrixfact05,bayesmatrix,cofirank}, or probabilistic latent model \cite{latent04,problatent}. However, as noted, numeric ratings can be highly varied even when preferences are shared. Pairwise preference based ranking methods can effectively address the limitations of rating based methods. Furthermore, numerical ratings can always be transformed into pairwise comparisons. Salimans et al.~\cite{salimans2012} use a bilinear model and do estimation in the Bayesian framework. Liu et al.~\cite{problatent} use the Bradley-Terry-Luce (BTL) Model. Rather than our low-rank setting, they characterize the similarity between different users by using a mixture model. Both methods are computationally inefficient. More important, all these methods fail to provide theoretical justifications of their algorithms. There are some theoretical works for learning a single ranking list from pairwise comparisons. Work by Jamieson and Nowak~\cite{JamNow11} seeks to exploit comparisons to significantly reduce the number of samples required to obtain a good estimate of an individual's utility function. Their method demonstrates that when the objects exist in a lower-dimensional space, then the number of queries required to learn the user's utility significantly decreases. One drawback of their approach is that the authors must assume that descriptors or features for the underlying objects are provided; which is not necessarily the case in all contexts. Negahban et al. \cite{NegOhSha12} propose the Rank Centrality algorithm and show rate optimal (up to log factors) error bounds of their algorithm under BTL model. They also provide theoretical analysis of penalized maximum likelihood estimator, which serves as an inspiration of our work. \comment{ As alluded to above there has been some work in understanding collaborative rankings and learning user preferences based on pairwise comparisons. Work by Jamieson and Nowak~\cite{JamNow11} seeks to exploit comparisons to significantly reduce the number of samples required to obtain a good estimate of an individual's utility function. Their method demonstrates that when the objects exist in a lower-dimensional space, then the number of queries required to learn the user's utility significantly decreases. One drawback of their approach is that the authors must assume that descriptors or features for the underlying objects are provided; which is not necessarily the case in all contexts. Nevertheless, their method is very natural when absolute ratings are not feasible to attain. In our setting, we wish to to rely on comparison based results from \emph{all} users to infer what items might be desirable to any given specific user. In that direction there have been recent algorithmic advancements~\cite{cofirank,param17}. These papers aim to optimize over the Normalized Discounted Cumulative Gain (NDCG). In order to do so, these papers must assume that the true underlying ratings are made available. Then, they wish to find rankings that are most consistent with the observed ratings (that is item $i$ is ranked higher than item $j$ if the item $i$ has a larger rating than item $j$.). These authors have shown that strong improvements over collaborative filtering can be made with respect to making accurate recommendations. However, underlying their method is the assumption that ratings can be observed and they exploit that information to optimize over the individual user rankings. Both papers also use implicit or explicit regularization to reduce over-fitting, either via a nuclear norm regularization~\cite{cofirank} as we do in this paper, or by using a nearest neighbor approach in providing recommendations~\cite{param17}. While these results present an important first step in understanding collaborative ranking, they do not provide theoretical justification or guarantees on the performance. Additionally, they assume that explicit ratings are available, which is not always the case. In contrast, we are focused on the setting where no explicit ratings are available to us and we wish to provide rigorous statistical guarantees on recovering the underlying preference parameters that model the choices of the users.} \paragraph{Our contributions} In this report, we present the first theoretical analysis of a collaborative ranking algorithm under a natural observation model. The algorithm itself is quite simple and falls into the framework of regularized $M$-estimator~\cite{Neg09}. We provide finite sample guarantees that hold with high probability on recovering the underlying preference matrix. Furthermore, the techniques outlined in the proof section our general and can be applied to a variety of sampling operators for matrix completion. For example, a simple modification of our proof yields a different class of results for the ``one-bit'' matrix completion problem~\cite{onebit}. In the following we present an explicit description of our model in Section~\ref{sec:model}. In Section~\ref{sec:estimate} we present the proposed estimation procedure that we wish to analyze. Finally, in Section~\ref{sec:main} we provide a statement of the main theorem followed by experiments in Section~\ref{sec:experiments}. Finally, in Section~\ref{sec:proofs} we present the proof. \comment{ Nevertheless, those methods demonstrate that while people's ratings might be inconsistent, their preferences (e.g. preferring item $a$ over item $b$) remain well behaved. There has been growing recent attention in understanding how to aggregate partial rankings from various users in a consistent manner. While there has been some theoretical advancements in this area, most of the work has focused on aggregating results into single rankings. This generalization of collaborative filtering yields the problem of collaborative ranking. This context is natural as in many settings a user might pick one items versus others: for example at the super market, when selecting a movie among a choice of movies, or when picking a webpage based on her search query. Other collaborative ranking approaches have proven to be very promising~\cite{param17,cofirank}, but theoretical justifications for their use have not been as explored while the standard matrix completion problem~\cite{CanRecCompletionCanRecCompletion} has received a significant amount of theoretical attention.} \comment{ Parameter recovery is important in this context because rather than being able to predict future queries, we wish to provide recommendations for future comparisons to be made. Thus, having parameter strengths that provide us some notion of how preferable an item is allows us to quickly make recommendations rather than simply predict the user's preference when given two items. } \comment{ In comparison to eigenrank, this approach explicitly builds a ranking and clustering simultaneously. While eigenrank relies on first finding ``close'' users and then performing a rank aggregation scheme similar to the one analyzed in Negahban, Oh, and Shah~\cite{NegOhSha12}, the method that we will explore performs the rank aggregation jointly without needing to first build a neighborhood structure. Implicitly the method does effectively find close neighbors and then aggregates the rankings; however, the user does not need to consider these aspects of the problem. Furthermore, we are able to provide strict theoretical guarantees on the performance of the algorithm. Again, showing that the sample complexity of our method matches state of the art results in the matrix completion setting. } \comment{ Other NDCG based approaches also exist~\cite{param17}, but those assume that true underlying ratings exist. In some situations, the user might not ever explicitly provide a relevance score. Thus, training with an NDCG objective might not be feasible. We do assume that there are true underlying ratings, but we work in a situation where these parameters are not directly accessible. Nevertheless, we are able to show consistent recovery of those ratings. Again, to the best of our knowledge, no known results exist that explicitly provide error guarantees for the behavior of comparison based recovery.} \comment{ The method that we introduce is related to work presented by Smola et. al.~\cite{smola}. They also use a factorization based approach. Those results focus on the algorithmic aspects of the problem and also do not provide theoretical guarantees on parameter recovery.} \paragraph{Notation:} For a positive integer $n$ we will let $[n] = \{1,2,\hdots,n\}$ be the set of integers from $1$ to $n$. For two matrices $A$, $B \in \mathbb{R}^{\dima \times \dimb}$ of commensurate dimensions, let $\tracer{A}{B} = \trace(A^T B)$ be the trace inner product. For a matrix $A \in \mathbb{R}^{\dima \times \dimb}$ let $A_{i,j}$ denote the entry in the $i^{th}$ row and $j^{th}$ column of $A$. Take $\svds{i}(A)$ to be the $i^{th}$ singular value of $A$ where $\svds{i}(A) \geq \svds{i+1}(A)$. Let $\opnorm{A} = \svds{1}(A)$, $\nucnorm{A} = \sum_{j=1}^{\min(\dima,\dimb)} \svds{j}(A)$ be the nuclear norm of $A$, i.e. the sum of the singular values of $A$, and $\frobnorm{A} = \sqrt{\tracer{A}{A}}=\sqrt{\sum_{j=1}^{\min(\dima,\dimb)} \svds{j}^2(A)}$ to be the Frobenius norm of $A$. Finally, we let $\infnorm{A} = \max_{i,j} |A_{i,j}|$ to be the elementwise infinity norm of the matrix $A$. \section{Problem Statement and Model} \label{sec:model} In this section we provide a precise description of the underlying statistical model as well as our problem. \subsection{Data and Observation Model} Recall that each user provides a collection of pairwise preferences for various items. We assume that the data are the form $(\design{i},\obs{i})$ where $\design{i} \in \mathbb{R}^{\dima \times \dimb}$. We assume that the $i^{th}$ piece of data is a query to user $\useri{i}$ asking if she prefers item $\itema{i}$ to item $\itemb{i}$. If she does, then $\obs{i}=1$, otherwise $\obs{i}=0$. In other words, $\obs{i} = 1$ if user $\useri{i}$ prefers item $\itema{i}$ to item $\itemb{i}$, otherwise $\obs{i}=0$. Let the underlying (unknown and unobservable) user preferences be encoded in the matrix $\Thetastar \in \mathbb{R}^{\dima \times \dimb}$ such that $\Thetastar_{k,j}$ is the score that user $k$ places on item $j$. We will also assume that $\frobnorm{\Thetastar} \leq 1$ to normalize the signal. For identifiability we assume that the sum of the rows of $\Thetastar$ is equal to zero. We must also assume that $\infnorm{\Thetastar} \leq \frac{\spiky}{\sqrt{\dima \dimb}}$. Similar assumptions are made in the matrix completion literature and is known to control the ``spikyness'' of the matrix. Both of these assumptions are discussed in the sequel. For compactness in notation we let $\design{i} = \sqrt{\dima \dimb} \stand{\useri{i}} (\stand{\itema{i}} - \stand{\itemb{i}})^T$ where $\stand{a}$ is the standard basis vector that takes on the value $1$ in the $a^{th}$ entry and zeros everywhere else. Taking the trace inner product between $\Thetastar$ and $\design{i}$ yields \begin{equation*} \tracer{\Thetastar}{\design{i}} = \sqrt{\dima \dimb} \ ( \Thetastar_{\useri{i},\itema{i}} - \Thetastar_{\useri{i},\itemb{i}} \ ) \end{equation*} and denotes the relative preference that user $\useri{i}$ has for item $\itema{i}$ versus $\itemb{i}$. Our observation model takes the form \begin{equation} \label{eq:model} \mathbb{P}(\obs{i}=1 | \itema{i}=l, \itemb{i}=j, \useri{i}=k) = \mylogit{\tracer{\Thetastar}{\design{i}}} \end{equation} The above is the standard Bradley-Terry-Luce model for pairwise comparisons. In full generality, one can also consider the Thurstone models for pairwise preferences. We shall take $\Thetastar$ to be low-rank or well approximate by a low-rank matrix. This is analogous to the matrix completion literature and models the fact that the underlying preferences are derived from latent low-dimensional factors. In this way, we can extract features on items and users without explicit domain knowledge. \paragraph{Discussion of assumptions:} In the above we assume that the $\ell_\infty$ norm of the matrix is bounded. This form of assumption is required for estimating the underlying parameters of the matrix and can be thought of as an incoherence requirement in order to ensure that the matrix itself is not orthogonal to the observation operator. For example, suppose that we have a matrix that is zeros everywhere except in one row where we have a single $+1$ and a single $-1$. In that case, we would never be able to recover those values from random samples without observing the entire matrix. Hence, the error bounds that we derive will include some dependency on the infinity norm of the matrix. If generalization error bounds are the desired outcome, then such requirements can be relaxed at the expense of slower error convergence guarantees and no guarantees on individual parameter recovery. Also noted above is the requirement that the sum of each of the rows of $\Thetastar$ must be equal to $0$. This assumption is natural owing to the fact that we can ever only observe the differences between the intrinsic item ratings. Hence, even if we could exactly observe all of those difference, the solution would not be unique up to linear offsets of each of the rows. We refer the reader to other work in matrix completion~\cite{CanRecCompletion,NegWaiCompletion} for a discussion of incoherence. \section{Estimation Procedure} \label{sec:estimate} We consider the following simple estimator for performing collaborating ranking. It is an example of a regularized $M$-estimator~\cite{Neg09}. \begin{equation} \label{estprocedure} \Thetahat = \argmin_{\Param \in \rspace} \underbrace{\frac{1}{n} \sum_{i=1}^n \log(1+\exp(\tracer{\Param}{\design{i}})) - \obs{i} \tracer{\Param}{\design{i}}}_{\Loss(\Param)} + \lambda \nucnorm{\Param}, \end{equation} where $\Loss(\Param)$ is the random loss function and \begin{equation*} \Omega = \{A \in \mathbb{R}^{\dima \times \dimb} \mid \infnorm{A} \leq \spiky, \text{ and $\forall j \in [\dima]$ we have $\sum_{k=1}^{\dimb} A_{j,k} = 0$} \} \end{equation*} This method is a convex optimization procedure, and very much related to the matrix completion problems studied in the literature. A few things to note about the constraint set presented above. While in practice, we do not impose the $\ell_\infty$ constraint, the theory requires us to impose the condition and an interesting line of work would be to remove such a constraint. A similar constraint appears in other matrix completion work~\cite{NegWaiCompletion}. As discussed above, the second condition is a fundamental one. It is required to guarantee identifiability in the problem even if infinite data were available. The method itself has a very simple interpretation. The random loss function encourages the recovered parameters to match the observations. That is, if $y_i = 1$ then we expect that $\Thetastar_{\useri{i},\itema{i}} > \Thetastar_{\useri{i},\itemb{i}}$. The second term is the nuclear norm and that encourages the underlying matrix $\Thetastar$ to be low-rank~\cite{CanRecCompletion}. \section{Main Results} \label{sec:main} In this section we present the main results of our paper, which demonstrates that we are able to recover the underlying parameters with very few total observations. The result is analogous to similar results presented for matrix completion~\cite{TsyCompletion,SewoongCompletion, NegWaiCompletion}, \begin{theos} \label{maintheorem} Under the described sampling model, let $d=(\dima + \dimb)/2$, assume $\numobs < d^2 \log d$, and take $\lambda \geq 32 \sqrt{\frac{\dimd \log \dimd}{\numobs}}$. Then, we have that the Frobenius norm of the error $\Delta = \Thetahat - \Thetastar$ satisfies \begin{equation*} \frobnorm{\Delta} \le c_1 \max \left ( \spiky,\frac{1}{\psi(2 \alpha)} \right ) \max \left \{\sqrt{\frac{r\dimd \log \dimd}{n}}, \left( \sqrt{\frac{r\dimd \log \dimd}{n}}\sum_{j=r+1}^{\min\{\dima, \dimb\}} \sigma_j(\Thetastar) \right )^{1/2} \right \} \end{equation*} with probability at least $1 - \frac{2}{d^2}$ for some universal constant $c_1$. \end{theos} The above result demonstrates that we can obtain consistent estimates of the parameters $\Thetastar$ using the convex program outlined in the previous section. Furthermore, the error bound behaves as a parametric error rate, that is the error decays as $\frac{1}{\numobs}$. The result also decomposes into two terms. The first is the penalty for estimating a rank $\rdim$ matrix and the second is the price we pay for estimating an approximately low-rank matrix $\Thetastar$ with a rank $r$ matrix. These results exactly match analogous results in the matrix completion literature barring one difference: there is also a dependency on the function $\psi$. However, this necessity is quite natural since if we are interested in parameter recovery, then it would be impossible to distinguish between extremely large parameters. Indeed, this observation is related to the problem of trying to measure the probability of a coin coming up heads when that probability is extremely close to one. Other results in matrix completion also discuss such a requirement as well as the influence of the spikyness parameter~\cite{TsyCompletion,onebit}. The proof of this result, for which we provide an outline in Section~\ref{sec:proofs}, follows similar lines as other results for matrix completion. \section{Experiments} \label{sec:experiments} Here we present simulation results to demonstrate the accuracy of the error rate behavior predicted by Theorem~\ref{maintheorem}. To make the results more clean, we consider the exact low rank case here, which means each individual user's preference vector is the linear combination of $r$ preference vectors. Then according to our main results, the empirical squared Frobenius norm error $\frobnorm{\Thetahat-\Thetastar}^2$ under our estimation procedure~\eqref{estprocedure} will be scaled as $\frac{rd\log d}{n}$. For all the experiments, we solved the convex program~\eqref{estprocedure} by using proximal gradient descent with step-sizes from~\cite{AgaNegWai} for fast convergence via our own implementation in R. \begin{figure}[h] \centering \begin{subfigure}{.5\textwidth} \centering \includegraphics[width=\linewidth]{\figdir/simulation1.eps} \caption{} \label{fig:sub1} \end{subfigure \begin{subfigure}{.5\textwidth} \centering \includegraphics[width=\linewidth]{\figdir/simulation2.eps} \caption{} \label{fig:sub2} \end{subfigure} \caption{\noindent Plots of squared Frobenius norm error $\frobnorm{\Thetahat-\Thetastar}^2$ when applying estimation procedure (\ref{estprocedure}) on the exact low rank matrix. Each curve corresponds to a different problem size $\dima=\dimb=d \in \{100,150,200,250\}$ with a fixed rank $r=4$. (a) Plots of Frobenius norm error against the raw sample size. As sample size increases, the error goes to zero. (b) Plots of the same Frobenius norm error against rescaled sample size $n/(rd\log d)$, all plots are aligned fairly well as expected by our theory.} \label{fig:simulation} \end{figure} In Figure 1 we report the results of four different problem sizes with equal user size $\dima$ and item size $\dimb$ and the fixed rank $r$, where $\dima=\dimb=d \in \{100,150,200,250\}$, $r=4$. For a given sample size d, we ran $T=10$ trials and computed the squared Frobenius norm error $\frobnorm{\Thetahat-\Thetastar}^2$ averaged over those trials. Panel (a) shows the plots of Frobenius norm error versus raw sample size. It shows the consistency of our estimation procedure because the Frobenius norm error goes to zero as sample size increases. And the curves shift to right as the problem dimension $d$ increases, matching with the intuition that larger matrices require more samples. In panel (b), we plot the simulation results versus the rescaled sample size $N=n/(rd\log d)$. Consistent with the prediction of Theorem \ref{maintheorem}, the error plots are aligned fairly well and decay at the rate of $1/N$ \section{Proof of Main Result} \label{sec:proofs} We now present a proof of the main result. We will use the machinery developed by Negahban and Wainwright~\cite{NegWaiCompletion} and establish a Restricted Strong Convexity (RSC) for our loss. The proof follows standard techniques, with some care when handling the new observation operator. \subsection{Proof of Theorem~\ref{maintheorem}} The key to establishing the RSC condition is to demonstrate that the error in the first order Taylor approximation of the loss is lower-bounded by some quadratic function. To that end we note that for $\Delta = \Param - \Thetastar$ and by the Taylor expansion we have that \begin{equation} \label{eq:secondordererror} \Loss(\Param) - \Loss(\Thetastar) - \tracer{\nabla \Loss(\Thetastar)}{\Delta} = \frac{1}{2 \numobs} \sum_{i=1}^\numobs \psi \left ( \tracer{\Thetastar}{\design{i}} + s \tracer{\Delta}{\design{i}} \right ) \left (\tracer{\Delta}{\design{i}} \right )^2, \end{equation} where $s \in [0,1]$ and \begin{equation*} \psi(x) = \frac{\exp(x)}{(1+\exp(x))^2}. \end{equation*} Now, we may apply the fact that both $\infnorm{\Thetahat}$, $\infnorm{\Thetastar} \leq \spiky/\sqrt{\dima \dimb}$ and that $\psi(x)$ is symmetric and decreases as $x$ increases to obtain that equation~\eqref{eq:secondordererror} is lower-bounded by: \begin{equation} \label{eq:lowerbound} \frac{1}{2 \numobs} \sum_{i=1}^\numobs \psi \left( 2 \spiky \right ) \left (\tracer{\Delta}{\design{i}} \right )^2 \end{equation} Therefore, it suffices to prove a lower-bound on $\frac{1}{2 \numobs} \left (\tracer{\Delta}{\design{i}} \right )^2$ for all possible vectors $\Delta$. For that, we present the following lemma. \begin{lems} \label{RSC} For $\infnorm{\Theta} \le r_3 := \frac{2\spiky}{\sqrt{\dima \dimb}}$, $d=(\dima + \dimb) /2$, and $\numobs < \dimd^2 \log \dimd$. When $\design{i}$ are i.i.d observations we have with probability greater than $1-2d^{-2^{18}}$ \begin{equation*} \frac{1}{\numobs} \sum_{i=1}^\numobs \left (\tracer{\Theta}{\design{i}} \right )^2 \ge \frac{1}{3}\frobnorm{\Theta}^2 ~~\text{~~for all $\Theta$ in $\setA$} \end{equation*} where \begin{equation*} \mathcal{A} = \left \{\Theta \in \mathbb{R}^{\dima \times \dimb} \mid \infnorm{\Theta} \leq r_3, \frobnorm{\Theta}^2 \ge 128 \spiky \sqrt{\frac{d \log d}{n}}\nucnorm{\Theta} \text{ and $\forall j \in [\dima]$ we have $\sum_{k=1}^{\dimb} \Theta_{j,k} = 0$} \right \} \end{equation*} \end{lems} Another key element for establishing the error is the following upper-bound on the operator norm of a random matrix. \begin{lems} \label{Lem:crosstermbound} Consider the sampling model described above. Then for i.i.d. $(\noise_i,\design{i})$, where $|\noise_i| \leq \gamma$ and $\mathbb{E}[\xi_i | \design{i}] = 0$ we have that \begin{equation*} \mathbb{P} \left ( \opnorm{\frac{1}{\numobs} \sum_{i=1}^\numobs \noise_i \design{i}} > 8\gamma \sqrt{\frac{ \dimd \log \dimd}{\numobs}} \right ) \le \frac{2}{\dimd^2}, \end{equation*} \end{lems} We these two ingredients in hand we may now prove the main result. The steps are a slight modification of the ones taken for standard matrix completion~\cite{NegWai11b}. By the optimality of $\Thetahat$ we have \begin{equation*} \Loss (\Thetahat) + \lambda \nucnorm{\Thetahat} \le \Loss (\Thetastar) + \lambda \nucnorm{\Thetastar} \end{equation*} Let $\Delta = \Thetahat - \Thetastar$, then \begin{equation*} \Loss(\Thetahat) - \Loss(\Thetastar) - \tracer{\nabla \Loss(\Thetastar)}{\Delta} \le - \tracer{\nabla \Loss(\Thetastar)}{\Delta} + \lambda \left ( \nucnorm{\Thetastar} - \nucnorm{\Thetahat} \right ) \end{equation*} By Taylor expansion, the left hand side is lower bounded by \begin{equation*} \Loss(\Thetahat) - \Loss(\Thetastar) - \tracer{\nabla \Loss(\Thetastar)}{\Delta} \ge \psi \left( 2\spiky \right ) \frac{1}{2\numobs} \sum_{i=1}^\numobs \left (\tracer{\Theta}{\design{i}} \right )^2 \end{equation*} H\"older's inequality between the nuclear norm and operator norm yields \begin{equation*} - \tracer{\nabla \Loss(\Thetastar)}{\Delta} \le \matsnorm{\nabla \Loss(\Thetastar)}{2} \nucnorm{\Delta} \end{equation*} By the triangle inequality $\nucnorm{\Thetastar} - \nucnorm{\Thetahat} \le \nucnorm{\Delta}$. If we choose $\lambda > 2\matsnorm{\nabla \Loss(\Thetastar)}{2}$, we have \begin{equation*} \Loss(\Thetahat) - \Loss(\Thetastar) - \tracer{\nabla \Loss(\Thetastar)}{\Delta} \le 2\lambda \nucnorm{\Delta} \end{equation*} Now, the random matrix $\nabla \Loss(\Thetastar)=\frac{1}{\numobs} \sum_{i=1}^{\numobs}\left ( \frac{\exp(\tracer{\design{i}}{\Delta})}{1+\exp(\tracer{\design{i}}{\Delta})} - y_i \right ) \design{i}$ and satisfies the conditions of Lemma \ref{Lem:crosstermbound} with $\gamma=2$, so we can take $\lambda=32 \sqrt{\frac{d \log d}{n}}$ From Lemma~1 of Negahban and Wainwright~\cite{NegWai11b}, $\Delta$ can be decomposed into $\Delta'+\Delta''$, where $\Delta'$ has rank less than $2r$ and $\Delta''$ satisfies \begin{equation*} \nucnorm{\Delta''} \le 3\nucnorm{\Delta'} + 4 \sum_{j=r+1}^{\min\{\dima, \dimb\}} \sigma_j(\Thetastar) \end{equation*} Then by the triangle inequality and $\nucnorm{\Delta'} \le \sqrt{2r} \frobnorm{\Delta'}$ \begin{equation} \label{nucbound} \nucnorm{\Delta} \le 4\nucnorm{\Delta'} + 4 \sum_{j=r+1}^{\min\{\dima, \dimb\}} \sigma_j(\Thetastar) \le 4\sqrt{2r} \frobnorm{\Delta} + 4 \sum_{j=r+1}^{\min\{\dima, \dimb\}} \sigma_j(\Thetastar) \end{equation} Now depending on whether $\Delta$ belongs to set $\setA$, we split into two cases. \\ \textbf{Case 1}: When $\Delta \notin \setA$, $\frobnorm{\Delta}^2 \le 128 \spiky \nucnorm{\Delta} \sqrt{\frac{\dimd \log \dimd}{n}}$. From Equation~\eqref{nucbound}, we get \begin{equation*} \frobnorm{\Delta} \le \spiky\max \left \{1024 \sqrt{\frac{r\dimd \log \dimd}{n}}, \left( 512 \sqrt{\frac{r\dimd \log \dimd}{n}}\sum_{j=r+1}^{\min\{\dima, \dimb\}} \sigma_j(\Thetastar) \right )^{1/2} \right \} \end{equation*} \textbf{Case 2}: Otherwise, from Lemma \ref{RSC}, with probability greater than $1-2d^{-2^{18}}$, $\Loss(\Thetahat) - \Loss(\Thetastar) - \tracer{\nabla \Loss(\Thetastar)}{\Delta} \ge \frac{\psi \left( 2\spiky \right )}{3} \frobnorm{\Delta}^2$. Therefore, the above equations yield \begin{equation*} \frobnorm{\Delta}^2 \; \leq \frac{192}{\psi(2 \spiky)} \sqrt{\frac{2r \dimd \log \dimd}{\numobs}} \nucnorm{\Delta}. \end{equation*} Now, performing similar calculations as above we have \begin{equation*} \frobnorm{\Delta} \le \frac{1}{\psi(2 \spiky)} \max \left \{1024 \sqrt{\frac{r\dimd \log \dimd}{n}}, \left( 512 \sqrt{\frac{r\dimd \log \dimd}{n}}\sum_{j=r+1}^{\min\{\dima, \dimb\}} \sigma_j(\Thetastar) \right )^{1/2} \right \}. \end{equation*} Combining the two displays above yields the desired result. \subsection{Proof of Lemma \ref{RSC}} We use a peeling argument~\cite{geer2000} as in Lemma~3 of \cite{NegWaiCompletion} to prove Lemma~\ref{RSC}. Before that, we first present the following lemma. \begin{lems} \label{deviations} Define the set \begin{equation*} \mathcal{B}(D) = \left \{ \Theta \in \mathbb{R}^{\dima \times \dimb} \mid \infnorm{\Theta} \le r_3, \frobnorm{\Theta} \le D, \nucnorm{\Theta} \le \frac{D^2}{128\spiky} \sqrt{\frac{n}{d \log d}} \right \} \end{equation*} and \begin{equation*} M(D) = \sup_{\Theta \in \mathcal{B}(D)} \left ( - \frac{1}{\numobs} \sum_{i=1}^\numobs \left (\tracer{\Theta}{\design{i}} \right )^2 + 2\frobnorm{\Theta}^2 \right ) \end{equation*} Then \begin{equation*} \mathbb{P} \left \{ M(D) \ge \frac{3}{2}D^2 \right \} \le \exp \{ -\frac{nD^4}{128\spiky^4}\} \end{equation*} \end{lems} Since for any $\Theta \in \setA$, \begin{equation*} \frobnorm{\Theta}^2 \ge 128 \spiky \sqrt{\frac{d \log d}{n}}\nucnorm{\Theta} \ge 128 \spiky \sqrt{\frac{d \log d}{n}}\frobnorm{\Theta} \end{equation*} then we have $\frobnorm{\Theta} \ge 128 \spiky \sqrt{\frac{d \log d}{n}} := \mu$. Consider the sets \begin{equation*} \mathcal{S}_{\ell} = \left \{ \Theta \in \mathbb{R}^{\dima \times \dimb} \mid \infnorm{\Theta} \le r_3, \beta^{\ell-1} \mu \le \frobnorm{\Theta} \le \beta^\ell \mu, \nucnorm{\Theta} \le \frac{D^2}{128\spiky} \sqrt{\frac{n}{d \log d}} \right \} \end{equation*} where $\beta=\sqrt{\frac{10}{9}}$ and $\ell=1,2,3\cdots$. \\ Suppose there exists $\Theta \in \setA$ such that $\frac{1}{\numobs} \sum_{i=1}^\numobs \left (\tracer{\Theta}{\design{i}} \right )^2 < \frac{1}{3}\frobnorm{\Theta}^2$. Since $\setA \subseteq\bigcup_{\ell=1}^{\infty} \mathcal{S}_{\ell} \subseteq\bigcup_{\ell=1}^{\infty} \mathcal{B}(\beta^\ell \mu)$, there is some $\ell$ such that $\Theta \in \mathcal{B}(\beta^\ell \mu)$ and \begin{equation*} - \frac{1}{\numobs} \sum_{i=1}^\numobs \left (\tracer{\Theta}{\design{i}} \right )^2 + 2\frobnorm{\Theta}^2 > \frac{5}{3} \frobnorm{\Theta}^2 \ge \frac{5}{3} \beta^{2\ell-2} \mu^2 = \frac{3}{2} (\beta^{\ell} \mu)^2 \end{equation*} Then by union bound, we have {\setlength\arraycolsep{2pt} \begin{eqnarray*} && \mathbb{P} \left \{ \exists ~~\Theta \in \setA, ~\frac{1}{\numobs} \sum_{i=1}^\numobs \left (\tracer{\Theta}{\design{i}} \right )^2 < \frac{1}{3}\frobnorm{\Theta}^2 \right \} \\ &\le& \sum_{\ell=1}^{\infty} \mathbb{P} \left \{ M(\beta^{\ell} \mu) > \frac{3}{2} (\beta^{\ell} \mu)^2 \right \}\\ &\le& \sum_{\ell=1}^{\infty} \exp \{ -\frac{n(\beta^{\ell} \mu)^4}{128\spiky^4}\} \\ &\le& \sum_{\ell=1}^{\infty} \exp \{ -\frac{ 4\ell (\beta-1) n\mu^4}{128\spiky^4}\} \\ &\le& 2 \exp \{ -\frac{ 4 (\beta-1) n\mu^4}{128\spiky^4}\} \\ &\le& 2 \exp \{ - 2^{18} \log d \} \end{eqnarray*}} where the second inequality is Lemma~\ref{deviations}, the third inequality is $\beta^\ell \ge \ell (\beta-1)$ and we use the fact that $n<d^2\log d$ for the last inequality. \subsection{Proof of Lemma~\ref{deviations}} Define \begin{equation*} Z = : \frac{1}{\dima \dimb} M(D) = \sup_{\Theta \in \mathcal{B}(D)}\frac{1}{n} \sum_{i=1}^\numobs \left [ \mathbb{E} \Big(\Theta_{k(i)l(i)}-\Theta_{k(i)j(i)} \Big)^2 -\Big(\Theta_{k(i)l(i)}-\Theta_{k(i)j(i)} \Big)^2 \right ] \end{equation*} Our goal will be to first show that $Z$ concentrates around its mean and then upper bound the expectation. We prove the concentration results via the bounded differences inequality~\cite{ledoux2001}; since $Z$ is a symmetric function of its arguments, it suffices to establish the bounded differences property with respect to the first coordinate. Suppose we have two samples of $(\useri{i}, \itema{i}, \itemb{i})_{i=1}^n$ that only differ at the first coordinate. {\setlength\arraycolsep{2pt} \begin{eqnarray*} Z- Z' &\le& \sup_{\Theta \in \mathcal{B(D)}} \Bigg[ \frac{1}{n}\sum_{i=1}^{n} (\Theta_{k'(i)l'(i)}-\Theta_{k'(i)j'(i)} \Big)^2 - \frac{1}{n}\sum_{i=1}^{n} \Big(\Theta_{k(i)l(i)}-\Theta_{k(i)j(i)} \Big)^2 \Bigg ] \\ &=& \sup_{\Theta \in \mathcal{B(D)}} \frac{1}{n} \Bigg( \Big(\Theta_{k'(1)l'(1)}-\Theta_{k'(1)j'(1)} \Big)^2 - \Big(\Theta_{k(1)l(1)}-\Theta_{k(1)j(1)} \Big)^2 \Bigg) \\ &\le& \frac{4r_3^2}{n} \end{eqnarray*}} Then by the bounded differences inequality, we have \begin{equation} \label{bdiff} \mathbb{P} \{ Z - \mathbb{E} Z \ge t\} \le \exp\{ -\frac{nt^2}{32r_3^4}\} \end{equation} In order to upper bound $\mathbb{E} Z$, we use a standard symmetrization argument. {\setlength\arraycolsep{2pt} \begin{eqnarray*} \mathbb{E} Z &=& \mathbb{E} \sup_{\Theta \in \mathcal{B(D)}}\frac{1}{n} \sum_{i=1}^n \left [ \mathbb{E} \Big(\Theta_{a(i)l(i)}-\Theta_{a(i)j(i)} \Big)^2 -\Big(\Theta_{a(i)l(i)}-\Theta_{a(i)j(i)} \Big)^2 \right ] \\ &\le& \mathbb{E} \sup_{\Theta \in \mathcal{B(D)}} \frac{2}{n} \sum_{i=1}^{n} \varepsilon_i \Big(\Theta_{a(i)l(i)}-\Theta_{a(i)j(i)} \Big)^2 \\ &=& \mathbb{E} \sup_{\Theta \in \mathcal{B(D)}} \frac{2}{n} \sum_{i=1}^{n} \varepsilon_i \tracer{e_{k(i)} (e_{l(i)}-e_{j(i)})^T}{\Theta}^2 \end{eqnarray*}} where $\varepsilon_i$ are i.i.d. Rademacher random variables. Since $|\Theta_{a(i)l(i)}-\Theta_{a(i)j(i)}| \le 2r_3$, we have by the Ledoux-Talagrand contraction inequality that \begin{equation*} \mathbb{E} \sup_{\Theta \in \mathcal{B(D)}} \frac{1}{n} \sum_{i=1}^{n} \varepsilon_i \tracer{e_{k(i)} (e_{l(i)}-e_{j(i)})^T}{\Theta}^2 \le 4 r_3 \mathbb{E} \sup_{\Theta \in \mathcal{B(D)}} \frac{1}{n} \sum_{i=1}^{n} \varepsilon_i\tracer{e_{k(i)} (e_{l(i)}-e_{j(i)})^T}{\Theta} \end{equation*} By an application of H\"older's inequality we have that \begin{equation} |\sum_{i=1}^{n} \varepsilon_i \tracer{e_{k(i)} (e_{l(i)}-e_{j(i)})^T}{\Theta} | \le \matsnorm{\sum_{i=1}^{n} \varepsilon_i e_{k(i)} (e_{l(i)}-e_{j(i)})^T }2 \nucnorm{\Theta} \end{equation} Let $W_i := \varepsilon_i e_{k(i)} (e_{l(i)}-e_{j(i)})^T$. $W_i$ is a zero-mean random matrix, and since \begin{equation*} \mathbb{E} [ W_i W_i^T] = \mathbb{E} [e_{k(i)} (e_{l(i)}-e_{j(i)})^T (e_{l(i)}-e_{j(i)}) e_{k(i)} ^T]= (2-\frac{2}{\dimb}) \frac{1}{\dima} \mathbf{I}_{\dima \times \dima} \end{equation*} and \begin{equation*} \mathbb{E} [ W_i^T W_i] = \mathbb{E} [(e_{l(i)}-e_{j(i)}) e_{k(i)} ^T e_{k(i)} (e_{l(i)}-e_{j(i)})^T ]= \frac{2}{\dimb} \mathbf{I}_{\dimb \times \dimb} - \frac{2}{\dimb^2} \1\1^T \end{equation*} we have \[ \sigma_i^2 = \max \{\matsnorm{\mathbb{E} [ W_i^T W_i]}2, \matsnorm{ \mathbb{E} [ W_i W_i^T] }2 \} \le\max \{ \frac{2}{\dimb}, (2-\frac{2}{\dimb}) \frac{1}{\dima} \} \le \frac{2}{\min\{\dima,\dimb\}} \] Notice $\matsnorm{W_i}2\le 2$, thus, Lemma~\ref{Ahlswede-Winter} yields the tail bound \begin{equation} \mathbb{P} \Big[ \matsnorm{\frac{1}{n} \sum_{i=1}^{n} \varepsilon_i e_{k(i)} (e_{l(i)}-e_{j(i)})^T}2 \ge t\Big] \le \dima \dimb \max \{ \exp(-\frac{nt^2 \min\{\dima,\dimb\} }{8}), \exp(-\frac{nt}{4})\} \end{equation} Set $t=\sqrt{\frac{16\log \dima\dimb}{n \min\{\dima,\dimb\}}}$, we obtain with probability greater that $1-\frac{1}{\dima \dimb}$, \begin{equation*} \matsnorm{\frac{1}{n} \sum_{i=1}^{n} \varepsilon_i e_{k(i)} (e_{l(i)}-e_{j(i)})^T}2 \le \sqrt{\frac{16\log \dima\dimb}{n \min\{\dima,\dimb\}}} \end{equation*} By the triangle inequality, $\matsnorm{\frac{1}{n} \sum_{i=1}^{n} \varepsilon_i e_{k(i)} (e_{l(i)}-e_{j(i)})^T}2 \le \matsnorm{\varepsilon_i e_{k(i)} (e_{l(i)}-e_{j(i)})^T}2 \le 2$ and the fact $n \le d^2\log d$ \begin{equation} \mathbb{E} \matsnorm{\frac{1}{n} \sum_{i=1}^{n} \varepsilon_i e_{k(i)} (e_{l(i)}-e_{j(i)})^T}2 \le \sqrt{\frac{16\log \dima\dimb}{n \min\{\dima,\dimb\}}} + \frac{2}{\dima \dimb} \le 8 \sqrt{\frac{\log \dima\dimb}{n \min\{\dima,\dimb\}}} \end{equation} Putting those bounds together we have \[ \mathbb{E} \sup_{\Theta \in \mathcal{B(D)}} \frac{1}{n} \sum_{i=1}^{n} \varepsilon_i \Big(\Theta_{a(i)l(i)}-\Theta_{a(i)j(i)} \Big)^2 \le \sup_{\Theta \in \mathcal{B(D)}} 32 r_3 \nucnorm{\Theta} \sqrt{\frac{\log \dima\dimb}{n \min\{\dima,\dimb\}}} \le \frac{D^2}{\dima \dimb} \] Plug it into \eqref{bdiff} and set $t = \frac{D^2}{2\dima \dimb}$, we get the result. \subsection{Ahlswede-Winter Matrix Bound} As in previous work~\cite{NegWaiCompletion} we also use a version of the Ahlswede-Winter concentration bound. We use a version due to Tropp~\cite{randommatrix}. \begin{lems}[Theorem~1.6~\cite{randommatrix}] \label{Ahlswede-Winter} Let $W_i$ be independent $\dima \times \dimb$ zero-mean random matrices such that $\matsnorm{W_i}2 \le M$, and define \[ \sigma_i^2 := \max \{\matsnorm{\mathbb{E} [ W_i^T W_i]}2, \matsnorm{\mathbb{E} [ W_i W_i^T] }2 \} \] as well as $\sigma^2 := \sum_{i=1}^{n} \sigma_i^2$. We have \begin{equation} \mathbb{P} \Big[ \matsnorm{\sum_{i=1}^{n} W_i }2 \ge t\Big] \le (\dima + \dimb) \max \{ \exp(-\frac{t^2}{4\sigma^2}), \exp(-\frac{t}{2M})\} \end{equation} \end{lems} \section{Discussion} In this paper we presented a theoretical justification for a ranking based collaborative filtering approach based on pairwise comparisons in contrast to other results that rely on knowing the underlying ratings. We provided the first convergence bounds for recovering the underlying user preferences of items and showed that those bounds are analogous to the ones originally developed for rating based matrix completion. The analysis here can also be extended do other observation models, for example to the ``one-bit'' matrix completion setting as well. However, that extension does not provide any additional insights beyond the analysis presented here. There remain a number of extensions for these methods including adaptive and active recommendations, skewed sampling distributions on the items, as well as different choice models. We leave such extensions for future work. \comment{\section{Acknowledgements} The authors would like to thank Sewoong Oh and Devavrat Shah for many helpful and inspiring conversations on rank aggregation.}
\section{Introduction} In a previous paper [Sch, 1998], a notion of smooth functions on the Cantor set was developed. Recall that a {\it totally disconnected} topological space has a basis of clopen sets. The Cantor set is such a space. In this paper, we attempt to construct smooth functions on the Stone-${\check {\rm C}}$ech compactification of the integers. In addition to being totally disconnected, this space is {\it extremely disconnected}, which means that the closure of every open set is clopen. We will be working with the $C^\star$-algebra of all bounded complex-valued functions (or sequences) on the integers ${\Bbb Z}$, under pointwise multiplication, with pointwise complex-conjugation for involution. We denote this algebra by $\ell^\infty({\Bbb Z})$. None of the theorems will use the additive and multiplicative structure of ${\Bbb Z}$, so that any countable discrete set can be substituted for ${\Bbb Z}$. However, when specific objects are constructed in the examples, we may make reference to the underlying set of integers. Recall that the integers ${\Bbb Z}$ consists of all whole numbers from $-\infty$ to $\infty$, whereas the natural numbers ${\Bbb N}$ contains only the whole numbers from $0$ to $\infty$. The norm on $\ell^\infty({\Bbb Z})$ is the sup-norm $\| \quad \|_\infty$, defined by $\| \varphi \|_\infty = \sup_{n \in {\Bbb Z}} |\varphi(n)|$. We will use the notation $c_0({\Bbb Z})$ for the complex-valued functions (sequences) on ${\Bbb Z}$ which vanish at infinity. \vskip\baselineskip \section{The Stone-${\check {\rm C}}$ech Compactification of $\Bbb Z$} We recall the standard definition of the Stone-${\check {\rm C}}$ech compactifiaction from [Roy, 1968]. Let $\bf F$ be the set of all real-valued functions from $\Bbb Z$ into the closed interval $I=[-1,1]$. (So $\bf F$ is the set of real-valued functions in the unit ball of $\ell^\infty({\Bbb Z})$.) Let $${\bf X} = \prod_{\bf F} I$$ be the $\bf F$-fold cartesian product of unit intervals $I$. By the Tychonoff theorem [Roy, 1968, Chapter 9, Theorem 19], this is a compact Hausdorff space. Let $$i:{\Bbb Z} \hookrightarrow {\bf X} {\rm \quad where \quad} i(n)\longmapsto\bigl\{f(n)\bigr\}_{f \in \bf F}$$ be the natural inclusion map. \noindent{\bf Definition 1.1. \ } We define the {\it Stone-${\check {\rm C}}$ech Compactification of ${\Bbb Z}$}, denoted by $\beta({\Bbb Z})$, to be the closure of the image $i({\Bbb Z})$ in ${\bf X}$. We let $C(\beta({\Bbb Z}))$ denote the continuous complex-valued functions on $\beta({\Bbb Z})$. See [Roy, 1968] for the basic properties of $\beta({\Bbb Z})$. \noindent{\bf Definition 1.2. \ } Let $n_0 \in \beta({\Bbb Z})$, and let $\{n_\alpha\}_{\alpha \in \Lambda}$ be a net converging to $n_0$, where $\Lambda$ is some directed set, and $n_\alpha \in {\Bbb Z}$ for each $\alpha \in \Lambda$. Let $\varphi \in \ell^\infty({\Bbb Z})$. Then {\it $\varphi(n_0)$} is defined as the limit of the net $\{\varphi(n_\alpha)\}_{\alpha \in \Lambda}$. In this way $\varphi$ defines a continuous function on $\beta({\Bbb Z})$, giving an isomorphism of commutative $C^\star$-algebras $\ell^\infty({\Bbb Z}) \cong C(\beta({\Bbb Z}))$. (The fact that $\varphi$ is continuous, and that the limit defining $\varphi(n_0)$ converges, can be verified directly from the definition of the Stone-${\check {\rm C}}$ech compactification above, using the fact that the real and imaginary parts of $\varphi/{\| \varphi \|_\infty}$ are functions in $\bf F$.) Let $p\in \ell^\infty({\Bbb Z})$ be a projection. In other words for each $n\in {\Bbb Z}$, $p(n) = 0$ or $p(n)=1$. Then $p$ also defines a projection on $\beta({\Bbb Z})$. In fact, if $n_0\in \beta({\Bbb Z})$ and $\{n_\alpha\}_{\alpha \in \Lambda}$ are as above, then $p(n_\alpha)$ is either eventually equal to $1$ (in the case $p(n_0)=1$) or eventually equal to $0$ (in the case $p(n_0)=0$). Thus each point $n_0 \in \beta({\Bbb Z})$ defines a family ${\cal F}_{n_0}$ of subsets of $\Bbb Z$, where $S \in {\cal F}_{n_0}$ if and only if the projection $p$ whose support is equal to $S$ satisfies $p(n_0)=1$. \noindent{\bf Definition 1.3. \ } A family of subsets $\cal F$ of ${\Bbb Z}$ is a {\it filter} on ${\Bbb Z}$ if it is closed under finite intersections $$S, T \in {\cal F} \Longrightarrow S\cap T \in {\cal F}\eqno{\rm (1.4a)} $$ and supersets $$S \in {\cal F} \quad{\rm and}\quad T \supset S \Longrightarrow T \in {\cal F},\eqno{\rm (1.4b)} $$ and if the empty set $\emptyset$ in {\it not} in ${\cal F}$. A filter $\cal U$ on ${\Bbb Z}$ is an {\it ultrafilter} on ${\Bbb Z}$ if $$S \subseteq {\Bbb Z} \Longrightarrow S \in {\cal F} \quad{\rm or}\quad S^{c} \in {\cal F}.\eqno{\rm (1.4c)} $$ The map $n_0 \in \beta({\Bbb Z}) \mapsto {\cal F}_{n_0}$ from the previous paragraph defines a map {\it ${\cal UF}$} from $\beta({\Bbb Z})$ to the set of ultrafilters on ${\Bbb Z}$. (Use the definition of $\beta({\Bbb Z})$ to check this.) A {\it principal filter} is of the form ${\cal U}_n = <S \subseteq {\Bbb Z} \mid n \in S>$ for some $n \in {\Bbb Z}$. Such a filter is also an ultrafilter. The image of ${\Bbb Z} \subseteq \beta({\Bbb Z})$ under the map ${\cal UF}$ is precisely the set of principal ultrafilters on ${\Bbb Z}$. We show that the map ${\cal UF}$ is an isomorphism by constructing an inverse map ${\cal FU}$. Let ${\cal U}$ be any ultrafilter on ${\Bbb Z}$. Define a directed set $\Lambda$ to be ${\cal U}$ with superset order. That is $\alpha \le \beta \Longleftrightarrow \beta \subseteq \alpha$. Thus smaller sets are \lq\lq bigger\rq\rq\ in this order. In the case of the principal ultrafilter $\Lambda={\cal U}_n$, the singleton $\{n\}$ is the biggest element. In general, $\Lambda$ has a biggest element if and only if it comes from a principal ultrafilter. Next, for $\alpha\in \Lambda$ choose any $n_\alpha\in \alpha \subseteq {\Bbb Z}$. We show that $\{ n_\alpha \}_{\alpha \in \Lambda}$ converges to an element of $\beta({\Bbb Z})$. Let $\varphi \in {\bf F}$, where ${\bf F}$ is the defining family of functions for $\beta({\Bbb Z})$ from Definition 1.1. We wish to show that $\{ \varphi(n_\alpha) \}_{\alpha \in \Lambda}$ converges. Define \begin{equation} \varphi_{+}(n) = \begin{cases} \varphi(n) &\text{if $\varphi(n)\ge 0$;}\\ 0 &\text{otherwise.} \end{cases} \nonumber \end{equation} and \begin{equation} \varphi_{-}(n) = \begin{cases} -\varphi(n) &\text{if $\varphi(n)\le 0$;}\\ 0 &\text{otherwise.} \end{cases} \nonumber \end{equation} Then $\varphi = \varphi_{+} - \varphi_{-}$, and $\varphi_{+}, \varphi_{-} \in {\bf F}$. It suffices to show that $\{ \varphi_{+} (n_\alpha) \}_{\alpha \in \Lambda}$ and $\{ \varphi_{-} (n_\alpha) \}_{\alpha \in \Lambda}$ each converge. So without loss of generality, we take $\varphi$ with range in the unit interval $[0,1]$. If $\varphi$ were a projection $p$, we would be done. Simply let $S$ be the support of $p$. If $S \in {\cal U} = \Lambda$, then $\{ p (n_\alpha) \}_{\alpha \in \Lambda}$ is eventually $1$ (when $\alpha \ge S$). Otherwise, it is eventually $0$, and in either case they converge. We proceed by writing $\varphi$ as an infinite series of projections. Define the set of integers $$S_{1\over 2} = \{ n \in {\Bbb Z} \mid {1\over 2} \le \varphi(n) \le 1 \}.$$ Let $p_{1\over 2}$ be the projection corresponding to $S_{1\over 2}$. Then $\varphi - {1\over 2}p_{1\over 2}$ has its range in the interval $[0,{1\over 2}]$. We repeat the process to get a new function with range in the interval $[0,{1\over 4}]$, etc, until we get: $$\varphi = {1\over 2}p_{1\over 2} + {1\over 4}p_{1\over 4} + \dots = \sum_{q=1}^{\infty} {1\over 2^q} p_q, \eqno(1.5)$$ an infinite series that converges absolutely (and geometrically fast) in sup norm to $\varphi$. To see that $\{ \varphi (n_\alpha) \}_{\alpha \in \Lambda}$ converges, now use a standard series argument. Let $\epsilon>0$ be given, and find $N$ large enough so that the sup norm of the tail $$ \biggl{\Vert} \sum_{q=N}^{\infty} {1\over 2^q} p_q \biggr{\Vert}_{\infty}$$ is less than $\epsilon$. Then find $\alpha \in \Lambda$ sufficiently large so that $\alpha \subseteq S_{1/ 2^q}$ or $\alpha \subseteq S^c_{1/ 2^q}$ for each $q=1,\dots N$. (One could first find an $\alpha$ that works for each $S_{1/ 2^q}$ separately, using the ultrafilter condition (1.4c). Then, find an $\alpha$ that works for all $S_{1/ 2^q}, q=1,\dots N$ using the finite intersection property (1.4a).) For any $\beta \in \Lambda$ beyond this $\alpha$, the first $N$ projections in (1.5) have settled down to their final value (either $0$ or $1$) on the net $\{ n_\alpha \}_{\alpha \in \Lambda}$. Thus the net of real numbers $\{ \varphi (n_\alpha) \}_{\alpha \in \Lambda}$ converges. Since $\{ n_\alpha \}_{\alpha \in \Lambda}$ was an arbitrary net from an arbitrary ultrafilter on ${\Bbb Z}$, this shows that ${\cal FU}$ is a well-defined map from the ultrafilters on ${\Bbb Z}$ into $\beta({\Bbb Z})$. One easily checks that the compositions ${\cal UF} \circ {\cal FU}$ and ${\cal FU} \circ {\cal UF}$ are identity maps, and so we have proved: \noindent{\bf Proposition 1.6. \ } The map ${\cal UF}$ is an isomorphism of the Stone-${\check {\rm C}}$ech compactification $\beta({\Bbb Z})$ with the set of ultrafilters on ${\Bbb Z}$. Under this map, a point $n_0\in \beta({\Bbb Z})$ is taken to the ultrafilter of sets that any net of integers converging to $n_0$ is eventually in. \vskip\baselineskip \section{Definition of the Smooth Functions $\ell^{\infty \infty}({\Bbb Z})$ ( also denoted by $C^\infty(\beta({\Bbb Z}))$ )} \noindent{\bf Definition 2.1. \ } Define {\it $\ell^\infty_c({\Bbb Z})$} to be the finite span of projections in $\ell^\infty({\Bbb Z})$. This is a dense $\star$-subalgebra of $\ell^\infty({\Bbb Z})$, and plays an analogous role to $\ell^\infty({\Bbb Z})$ as $c_c(\Bbb Z)$, the compact (or finite) support functions on $\Bbb Z$, does to $c_0({\Bbb Z})$. (The series expansion (1.5) proves the density.) For $n_0 \in \beta({\Bbb Z})$, choose a net $\{ n_\alpha \}_{\alpha \in \Lambda}$ of integers converging to $n_0$. Define the {\it smooth functions on $\beta({\Bbb Z})$}, denoted by $\ell^{\infty \infty}({\Bbb Z})$ or $C^\infty(\beta({\Bbb Z}))$, to be those functions $\varphi$ in $\ell^\infty({\Bbb Z})$ that satisfy $$\lim_{\alpha \in \Lambda} n_\alpha^d \biggl(\varphi(n_\alpha) - \varphi(n_0)\biggr) = 0 \eqno {(2.2)} $$ for each $d=0, 1, 2, \dots$ and for each $n_0 \in \beta({\Bbb Z})$. This set of functions $\ell^{\infty \infty}({\Bbb Z})$ is a dense $\star$-subalgebra of $\ell^\infty({\Bbb Z})$, which contains $\ell^\infty_c({\Bbb Z})$, and plays an analogous role to $\ell^\infty({\Bbb Z})$ as ${\cal S}({\Bbb Z})$, the Schwartz functions on $\Bbb Z$, does to $c_0({\Bbb Z})$. If $p$ is a projection in $\ell^\infty({\Bbb Z})$, we noticed above that $p(n_0)=0$ or $p(n_0)=1$ for any $n_0 \in \beta({\Bbb Z})$. In either case, the quantity in parentheses in (2.2) eventually becomes $0$, so that $\varphi = p$ satisfies (2.2). This shows that $\ell^\infty_c({\Bbb Z}) \subseteq \ell^{\infty \infty}({\Bbb Z})$. \noindent{\bf Lemma 2.3. \ } The limit (2.2) holds independently of the choice of net $\{ n_\alpha \}_{\alpha \in \Lambda}$ converging to $n_0$. \noindent{\bf Proof:} Let $\epsilon > 0 $ be given and find a $\beta$ such that $ \bigl| n_\alpha^d \bigl( \psi(n_\alpha) - \psi(n_0) \bigr) \bigr| < \epsilon $ for $\alpha \ge \beta$. The set $S=\bigcup \{ n_\alpha \mid \alpha \ge \beta \}$ is in the ultrafilter associated with $n_0$ and $ \bigl| m^d \bigl( \psi(m) - \psi(n_0) \bigr) \bigr| < \epsilon $ for $m \in S$. If $\{ m_\alpha \}_{\alpha \in \Gamma}$ is another net tending to $n_0$, then it is eventually in $S$. So we have $ \bigl| m_\alpha^d \bigl( \psi(m_\alpha) - \psi(n_0) \bigr) \bigr| < \epsilon $ for $\alpha \ge \gamma$, for some $\gamma \in \Gamma$. {\bf QED} To see that $\ell^{\infty \infty}({\Bbb Z})$ is closed under products, let $\varphi$, $\psi \in \ell^{\infty \infty}({\Bbb Z})$. Then evaluate the quantity in parentheses in (2.2), namely the difference $$\varphi(n_\alpha) \psi(n_\alpha) - \varphi(n_0) \psi(n_0) = \biggl( \varphi(n_\alpha) - \varphi(n_0) \biggr) \psi(n_\alpha) + \varphi(n_0) \biggl( \psi(n_\alpha) - \psi(n_0) \biggr).$$ So the absolute value of the quantity in the limit (2.2) is $$\biggl| n_\alpha^d \biggl( \varphi\psi(n_\alpha)-\varphi\psi(n_0)\biggr)\biggr|\le \biggl| n_\alpha^d \biggl( \varphi(n_\alpha) - \varphi(n_0) \biggr) \biggr| \| \psi \|_\infty + \| \varphi \|_\infty \biggl| n_\alpha^d \biggl( \psi(n_\alpha) - \psi(n_0) \biggr) \biggr| $$ Clearly, this tends to zero as $n_\alpha$ tends to $n_0$. Next, we note that $\ell^{\infty \infty}({\Bbb Z})$ is actually bigger than $\ell^\infty_c({\Bbb Z})$. For example, any function in ${\cal S}({\Bbb Z})$ satisfies the limit (2.2), so $\ell^{\infty \infty}({\Bbb Z}) \supseteq \ell^\infty_c({\Bbb Z}) + {\cal S}({\Bbb Z})$. For $\varphi \in {\cal S}({\Bbb Z})$, note that $\varphi(n_0) = 0$ for any $n_0 \in \beta({\Bbb Z}) - {\Bbb Z}$. (Any nonprincipal ultrafilter eventually leaves every finite set.) Therefore $$\biggl| n_\alpha^d \biggl( \varphi(n_\alpha)-\varphi(n_0)\biggr)\biggr| = \biggl| n_\alpha^d \varphi(n_\alpha) \biggr| = 1/n_\alpha^2 \biggl| n_\alpha^{d+2} \varphi(n_\alpha) \biggr| \le 1/n_\alpha^2 \| \varphi \|_{d+2}, $$ where $\| \quad \|_{d+2}$ denotes the $d+2$th Schwartz seminorm on ${\cal S}({\Bbb Z})$. \noindent{\bf Proposition 2.4. \ } The inclusions $\ell^\infty_c({\Bbb Z}) + {\cal S}({\Bbb Z}) \subseteq \ell^{\infty \infty}({\Bbb Z}) \subseteq \ell^\infty({\Bbb Z})$ are proper. \noindent{\bf Proof:} The function $1\over{n^2 + 1}$ in $c_0({\Bbb Z}) - {\cal S}({\Bbb Z})$ does not satisfy (2.2), so the second inclusion is proper. The first inclusion is proper since $e^{i{\cal S}({\Bbb Z})} \subseteq \ell^{\infty \infty}({\Bbb Z})$. {\bf QED} Let $\varphi \in \ell^{\infty \infty}({\Bbb Z})$. We may write $\varphi$ (in fact any function in $\ell^\infty({\Bbb Z})$) uniquely in the form $$ \varphi = \sum_{q=1}^\infty c_q p_q, \eqno (2.5) $$ where the $p_q$'s are {\it pairwise disjoint projections}, and the $c_q$'s are {\it distinct} constants. \noindent{\bf Proposition 2.6. \ } At most finitely many projections in the series (2.5) have infinite support. \noindent{\bf Proof:} {\it We assume for a contradiction that there are infinitely many projections in (2.5), each having infinite support}. Note that the coefficients of the infinite projections must have an accumulation point. By dropping to a subsequence, and getting rid of all the finite projections, we may assume that $c_q \rightarrow c_0 \in {\Bbb C}$ as $q \rightarrow \infty$. (Multiply $\varphi$ by an appropriate projection, and renumber the $c_q$'s.) Let $S_q\subseteq {\Bbb Z}$ denote the support of the projection $p_q$. Then each set $S_q$ is infinite. Define a decreasing sequence of infinite subsets of $\Bbb Z$ by the disjoint unions $$U_n = \bigcup_{q \ge n} \biggl( S_q \bigcap \biggl\{ m \in {\Bbb Z}\,\, \biggl| \,\, m \ge {1\over { | c_q - c_0 | }} \biggr\} \biggr) $$ for $n = 1, 2, \dots $. Since finite intersections of the sets $U_n$'s are non-empty, there exists an ultrafilter $\cal U$ for which $U_n \in {\cal U}$ for each $n$. Let $\{ n_\alpha \}_{\alpha \in \cal U}$ be a net of integers that is eventually in this ultrafilter. This net must therefore also eventually be in each of the sets $U_n$. By construction, the point $n_0 \in {\cal FU}({\cal U}) \in \beta({\Bbb Z})$ that $\{ n_\alpha \}_{\alpha \in {\cal U}}$ converges to must satisfy $\varphi(n_0)= c_0$. If $n_\alpha \in S_q \bigcap \{ m \in {\Bbb Z} \mid m \ge {1\over |c_q - c_0|} \}$, then \begin{eqnarray} \biggl| n_\alpha^d \biggl( \psi(n_\alpha) - \psi(n_0) \biggr) \biggr| & = \biggl| n_\alpha^d \biggl( c_q - c_0 \biggr) \biggr| \qquad \qquad {\rm since} \quad n_\alpha \in S_q \nonumber \\ & \ge \bigl| n_\alpha^d \times {1\over {n_\alpha}} \bigr| = \bigl| n_\alpha^{d-1} \bigr| \ge 1, \nonumber \end{eqnarray} contradicting the fact that $\varphi$ must satisfy (2.2) for $d \ge 2$ at the point $n_0 \in \beta({\Bbb Z})$ we constructed. {\bf QED} It follows from Proposition 2.6 that for any $\varphi \in \ell^{\infty \infty}({\Bbb Z})$, we may substract an element of $\ell^\infty_c({\Bbb Z})$ to force the expansion (2.5) to have only projections of finite support. \noindent{\bf Remark 2.7. \ } Note that ${\cal S}({\Bbb Z})$ is an ideal in $\ell^{\infty \infty}({\Bbb Z})$ and $\ell^\infty({\Bbb Z})$. The closure ${\overline {{\cal S}({\Bbb Z})}}^{\| \,\,\,\, \|_\infty}$ is equal to $c_0({\Bbb Z})$, so ${\cal S}({\Bbb Z})$ is not dense in $\ell^\infty({\Bbb Z})$. The algebra $\ell^\infty_c({\Bbb Z})$, being unital, is {\it not} an ideal in either algebra $\ell^{\infty \infty}({\Bbb Z})$ or $\ell^\infty({\Bbb Z})$, and for the same reason $\ell^{\infty \infty}({\Bbb Z})$ is not an ideal in $\ell^\infty({\Bbb Z})$. This is in contrast to the case $c_c({\Bbb Z}) \subseteq {\cal S}({\Bbb Z}) \subseteq c_0({\Bbb Z})$, where every algebra is a dense ideal in every algebra above it. \vskip\baselineskip \vskip\baselineskip \section{References} \smallskip \smallskip \noindent[{\bf Roy, 1968}] \, H.L. Royden, {\it Real Analysis}, second edition, Macmillan Publishing Co., Inc., New York, 1968. \noindent[{\bf Sch, 1998}] \, L. B. Schweitzer, {\it $C^{\infty}$ functions on the cantor set, and a smooth $m$-convex Fr\'echet subalgebra of $O_2$}, Pacific J. Math. {\bf 184(2)} (1998), 349-365. \vskip\baselineskip \noindent{Email: <EMAIL>. Web Page: http://www.svpal.org/$\thicksim$lsch/Math.} \end{document}
\section{Introduction} The study of globular clusters (GCs) offers a rich platform for the validation of stellar evolution theories and the formulation of dynamical evolution scenarios of stellar systems.~In particular, recent discoveries \citep[e.g.][]{fer12,knigge09} have prompted Blue Straggler Stars (BSSs) as key elements of the puzzle, given their tight relation with these two fields.~BSSs are empirically defined as stars that appear bluer and brighter than turn-off stars in the color-magnitude diagram (CMD) of a GC. Although their definite physical nature is still a matter of debate, the most popular formation channels for BSSs are the ones explaining them as products of mass-transfer/mergers in binary stellar systems \citep{mccrea64} and as the results of direct stellar collisions \citep{hillsday76}.~Both are strongly dependent on the local environment's dynamical state, and, therefore, BSSs can be used as tracers of past dynamical events as well as of the current dynamical state, given their strong response to two-body relaxation effects due to their relatively high stellar masses.~\cite{fer12} proposed that the current BSS radial density profile can be used as a ``dynamical clock" to estimate the dynamical age of a GC, and continued to classify a large set of GCs in three different classes according to their dynamical state.~The dynamically oldest category contains, among others, GCs M\,30 and NGC\,362, which have been studied in this context more in detail by \citet[hereafter F09]{fer09} and \citet[hereafter D13]{dal13}.~These studies have found that the BSS population is indeed consistent with dynamically old GCs.~Furthermore, their CMDs reveal well defined double BSS sequences, which the authors claim are the result of single short-lived dynamical events, such as core-collapse.~It is believed that during such core contraction the stellar collision rate would become enhanced, producing the bluer BSS sequence in the CMD, while the boosted binary interaction rate would lead to enhanced Roche-lobe overflows, thereby producing the redder BSS sequence.~It is, therefore, clear that observational properties of BSSs can provide valuable information for the understating of the dynamical evolution of GCs.~Here, we present a study based principally on high-quality HST photometry of the inner BSS population in the poorly studied Galactic GC NGC\,1261, which contains evidence for a very particular dynamical history, making it a similar, yet unique case among other GCs with well defined BSS features, such as M\,30 and NGC\,362. \section{Data Description} The inner region photometric catalog comes from the HST/ACS Galactic Globular Cluster Survey \citep{sar07}.~It consists of $\sim$30 min.~exposures in the F606W ($\sim\!V$) and F814W ($\sim\!I$) bands for the central $3.4\arcmin\times3.4\arcmin$ field of NGC\,1261.~The photometry was corrected to account for updated HST/ACS WFC zero points and calibrated in the Vega photometric system.~The catalog provides high quality photometry down to $\sim\!6$ mag below the main-sequence turn-off. Additionally, we performed PSF photometry using the DoPHOT software package \citep{sch93, alo12} on HST/WFC3 data taken with the F336W ($\sim\!U$) band, available from the Hubble Legacy Archive (PI: Piotto, Proposal ID: 13297).~The photometry was calibrated using Stetson standards \citep{ste00} from the NGC\,1261 field. The astrometry was refined with the HST/ACS optical catalog using bright isolated stars, after which we reach a median accuracy of $\sim\!0.002\arcsec$ between the optical and F336W-band catalogs.~We also use wide-field photometry from the catalog published by \cite{kra10}, built from observations at the 1.3-m Warsaw telescope at Las Campanas Observatory, using a set of $UBVI$ filters and a $14\arcmin\times14\arcmin$ field of view. The photometry is calibrated to \cite{ste00} and the median error is $\leqslant$0.04 mag for all filters and colors down to $V=20$ mag.~The complete description of their data reduction and photometric calibration can be found in \cite{kra10}. \section{The central BSS population in NGC\,1261} BSSs are selected through their position in a CMD. Having three filters, we can use the additional color information to remove contaminants. First, we cross-match the optical ACS catalog with the F336W-band catalog and keep matched sources with separations $<$ 0.02\arcsec\ ($\sim\!0.5$ pix) and reported errors\footnote{Taken from the HST/ACS catalog. More information on the errors is found on the catalog's README file.} $<$ 0.03 mag in the F606W and F814W filter.~This results in $\sim\!25000$ sources in a $\sim\!2.7\arcmin\!\times\!2.7\arcmin$ field centered on NGC\,1261 (see Figure~\ref{cmd1}).~We use a F814W\,$<$\,19.5 mag limit for the BSS selection criteria from \cite{lei11}, who define the BSS region based on magnitude and color cuts in the (F606W-F814W)\,vs.\,F814W CMD, shown by the polygon in Figure~\ref{cmd1}.~All stars inside the region are considered to be BSS candidates, and used in the following analysis (unless removed from the sample; see further down in the text). \begin{figure}[t!] \centering \includegraphics[width=9cm]{fig1.pdf} \caption{\small{(F606W-F814W)\,vs.\,F814W CMD of the NGC\,1261 inner region. All detections come from the final matched HST catalog.~The R-BSS (red triangles), B-BSS (blue squares) and eB-BSS (black circles; filled circle is explained in Fig~\ref{Ucmd}) sub-samples are marked; crosses mark the rest of the BSS sample (except for diagonal crosses, see Fig~\ref{Ucmd}).~Pentagon symbols mark special cases, see Fig~\ref{Ucmd} .~The inset panel shows the distribution of BSS perpendicular distances from the best-fit line to the B-BSS sequence (shown as a dashed line) in mag units.~The solid red line shows a non-parametric Epanech\-nikov-kernel probability density estimate with 90\% confidence limits represented by the dotted pink lines.} } \label{cmd1} \end{figure} \begin{figure}[t!] \centering \includegraphics[width=9cm]{fig2.pdf} \caption{\small{(F336W-F814W)\,vs.\,F336W CMD of the NGC\,1261 inner region.~Filled and $\times$-shape symbols indicate BSS candidates that were rejected from our BSS sample.~Pentagon symbols mark BSS candidates that appear slightly faint in the F336W band, relative to their parent subsamples.} } \label{Ucmd} \end{figure} We identify two prominent BSS sequences in the CMD, each containing $\sim\!20$ BSSs (see inset panel).~We label them the blue-sequence-BSSs (B-BSSs; shown as blue squares) and the red-sequence-BSSs (R-BSSs, red triangles).~Similar to M\,30 and NGC\,362\footnote{Note that in those studies, the HST F555W filter is used as a V band proxy, while we use the F606W filter.}, the B-BSS sequence is narrower and better defined, while the R-BSS sequence is dispersed towards redder colours.~We note the appearance of a group of BSSs lying bluer from the B-BSS sequence in the CMD. Since they cannot be directly associated with the B-BSS sequence, as they are clearly separated in the diagram, we choose to label them as extremely-blue-BSSs (eB-BSSs, black circles) and we later discuss their likelihood of being associated to a particular BSS population.~This bluer component has neither been observed in M\,30 nor in NGC\,362, and, if confirmed real, requires a given BSS formation scenario, particular to the history of NGC\,1261, that is not present in the other two GCs. We now introduce the F336W-band photometry for contaminant detection.~We show in Figure~\ref{Ucmd} the (F336W-F814W)\,vs.\,F336W CMD and use the same symbols for the BSSs as in Figure~\ref{cmd1}.~Three BSS candidates plus another star from the eB-BSS group cannot be distinguished from $normal$ MS/SGB stars based on their location in the (F336W-F814W)\,vs.\,F336W CMD (shown as diagonal crosses and a filled circle, respectively), and hence we exclude them from the full BSS sample and subsequent analysis.~It is worth noting that two R-BSSs and another two eB-BSSs in Figure~\ref{Ucmd} (additionally marked with a pentagon) also show significantly redder (F336W-F814W) colors than the rest of their groups.~This result may point into photometric variability (since the F336W and F814W-band images on HST were taken $\sim\!7$ years apart, while the F606W and F814W-band exposures are less than an hour apart).~Indeed W\,UMa eclipsing binary systems are frequent among BSSs and have been detected in the double BSS sequences of NGC 362 (D13) and M\,30 (F09).~In particular, the W\,UMa BSSs detected in NGC\,362 were found to show a typical variability of 0.3 mag in all F390W, F555W and F814W bands (see Figure~10 in D13) which could account for the deviations found in (F336W-F814W) colors for these BSSs in NGC\,1261.~Another explanation could be the blend of cooler stars, which would explain the apparent missing flux in the F336W band.~However, the fact that the source centroids were requested to match within 0.02\arcsec\ ($\sim$0.2 FWHM in the optical) in the optical and near-ultraviolet, along with the conservative 0.03 mag error limit in the optical bands, points rather towards a well-fitted single PSF detection.~We compare now the identified sub-samples with isochrones from stellar collisional models of \cite{sil09}.~The models have a metallicity of $Z\!=\!0.001$ ([Fe/H]~$\!=\!-1.27$) and solar-scaled chemical composition ([$\alpha$/Fe]~$\!=\!0$).~Pairs of stars with masses between 0.4 and 0.8\,$M_\odot$ are collided using the MMAS software package \citep{lom02}.~The parent stars are assumed to be non-rotating, and 10 Gyr old at the time of the collision.~The collision products are evolved using the Monash stellar evolution code, as described in \cite{sil09}.~The time of the collision was taken to be $t\!=\!0$, and isochrones of various ages were calculated by interpolating along the tracks to determine the stellar properties (effective temperature, luminosity, etc.) at ages between 0.2 and 5 Gyr.~For the isochrones, we used collision products of the following stellar mass combinations: $0.4\!+\!0.4, 0.4\!+\!0.5, 0.4\!+\!0.6, 0.5\!+\!0.6, 0.6\!+\!0.6$, and $0.8\!+\!0.8\,M_\odot$.~We have adopted the distance modulus and reddening values of NGC\,1261 from \cite{dot10} and augmented the distance modulus by 0.25 mag in order to match the low-mass end of the collisional isochrone to the location of the main-sequence fiducial line in the CMD\footnote{This is mostly caused by the different metallicity/alpha enhancement between our models and the cluster, which has a metallicity closer to [Fe/H]$\sim-$1.35...$-$1.38 according to the latest studies \citep{kra10,pau10,dot10}, and a large $\alpha$-enhancement as it is expected for GCs in the halo \citep{pritzl05,woodley10} .}.~We find that the location of the B-BSSs and eB-BSSs can be reproduced by 2 Gyr and 200 Myr old isochrones of the stellar collisional models, respectively, as seen in Figure~\ref{cmd_iso}.~We note that the agreement of the model with the B-BSSs is remarkably good.~The eB-BSS sub-sample, although much less populated, is also somewhat consistent with the collisional 0.2\,Gyr isochrone model.~This enlarges the likelihood of this BSS feature being a real distinct population.~The R-BSSs need further explanation.~F09 and D13 found for M\,30 and NGC\,362 that the lower bound of the R-BSSs could be fairly well bracketed by the zero-age main-sequence (ZAMS) shifted by 0.75 mag towards brighter luminosities in the $V$ (F555W in their case) band, which approximately indicates the region populated by mass-transfer binaries, as predicted by \cite{tia06}.~In our case we use the 0.25 Gyr old Dartmouth isochrones \citep{dot08}, and find that the R-BSS sequence can only be reproduced by a region bracketed by the 0.25 Gyr old isochrone shifted by 0.45 mag and 0.75 mag to brighter F606W-band luminosities (grey region in the Figure). This shift is less than that required for M\,30 and NGC\,362 , and we note that a small difference is expected due to slightly different HST filters (F606W vs. F555W) as a $V$ band proxy. However, the prediction by \citeauthor{tia06} is based only on Case A (main-sequence donor) mass-transfer models.~\cite{lu10} showed that case B (red-giant donor) mass-transfer products lie indeed in a bluer region than the ZAMS+0.75 mag boundary.~Our current understanding of binary mass transfer is limited and, hence, our observations could help putting constraints on future binary stellar evolution models. \begin{figure}[t!] \centering \includegraphics[width=9cm]{fig3.pdf} \caption{\small{(F606W-F814W)\,vs.\,F606W CMD of the NGC\,1261 inner region with overplotted collisional isochrones of 2 Gyr and 200 Myr old, up to 1.3 $M_\odot$ and 1.6 $M_\odot$, respectively.~The grey band shows the zero-age main sequence isochrone, shifted by 0.45 and 0.75\,mag to brighter luminosities, to match the R-BSS sequence.~The red line is a Dartmouth isochrone with cluster parameters adopted from \cite{dot10}.~The representative photometric errors as obtained in the HST/ACS catalog are plotted in red bars.} } \label{cmd_iso} \end{figure} \section{Dynamical State of NGC\,1261} F09 and D13 have demonstrated that M\,30 and NGC\,362 show signs of being in an advanced state of dynamical evolution, as revealed by their centrally segregated BSS radial profiles \citep[which puts them in the {\sc Family~III} group in][]{fer12} and by their centrally peaked radial stellar density profiles.~Likewise, we plot in Figure~\ref{BSS_norm_prof} the normalized, cumulative BSS radial density profile for all BSS candidates and for the R-BSS and B-BSS sequences in NGC\,1261 out to 3.8 core radii, i.e.~the extent of the HST observations.~As expected, the whole BSS sample, as well as each BSS sequence, are more centrally concentrated than the reference SGB population\footnote{We choose SGB stars from the CMD inside the 19.2$<$F606W$<$19.5 mag range.}.~However, contrary to what was found in M\,30 and NGC\,362, we find the B-BSS component more centrally concentrated than the R-BSS component.~A K-S test shows that the null-hypothesis of these stars being drawn from the same radial distribution has a p-value of 0.33, which implies a non-negligible likelihood for common parent distributions.~Nevertheless, the difference between these profiles and the ones for M\,30 and NGC\,362 is still significant and motivates a discussion on possible qualitative differences of BSS formation history in NGC\,1261 versus the other GCs.~We do not find any B-BSSs inside $\sim\!0.5\,r_c$ (or about 10\arcsec), in agreement with F09 and D13, who as well report no B-BSSs within the inner 5-6\arcsec ($\sim\!1.5\,r_c$ and $\sim\!0.5\,r_c$ respectively).~These authors suggest that dynamical kicks are responsible for clearing the innermost region of any B-BSSs.~The detailed process is, however, unknown as accurate dynamical models are numerically expensive and, thus, still lacking. \begin{figure}[t!] \centering \includegraphics[width=9cm]{fig4.pdf} \caption{\small{Normalized cumulative radial distribution of BSSs in the central 3.8$\,r_c$ of NGC\,1261.~The full BSS sample (solid black line), B-BSSs (long-dashed blue line) and R-BSSs (short-dashed red line) are plotted.~The corresponding cumulative distribution of the reference SGB population is shown as the shaded region.~We use the centre of gravity coordinates RA~$\!=\!$~03h\,12m\,16.21s, DEC~$\!=\!-55^{\rm o}\,12\arcmin\,58.4\arcsec$ given by \cite{gol10}, and $r_c=0.35'$ \citep{har96}.}} \label{BSS_norm_prof} \end{figure} The results from Figure~\ref{BSS_norm_prof} alone cannot be used to suggest an advanced dynamical state in NGC\,1261, as the central concentration of BSSs is now known to be ubiquitous among all studied GCs.~A more complete understanding can be obtained by looking at the BSS radial profile at larger cluster-centric distances, as the radial distance of the normalized BSS density minimum will depend on the cluster's dynamical age.~\cite{fer12} showed that as a GC evolves dynamically the BSS radial density profile takes the form of a central high maximum and a secondary peak at larger radii with a minimum in between. In order to follow this approach we include the wide-field photometry catalog in our analysis and carefully merge it \footnote{The resulting merged SGB radial profile distribution is checked to be smooth and continuous, therefore, ruling out severe completeness issues for the BSSs, which are in a similar magnitude range.~The merging point is chosen at 3.8 $r_c$.} with the inner-region ACS catalog, providing us with a sampling of NGC\,1261 out to $r\!>\!30\,r_c$. We plot in Figure~\ref{BSS_rad_frac} the relative fraction of BSSs to SGBs\footnote{They were selected using a $19.2\!<\!V\!<\!19.5$ mag range cut. We use the ``$V$ $ground$" magnitude on the ACS catalog in order to make them compatible with the wide-field gound-based photometry.} as a function of cluster-centric radius, assuming Poissonian noise for the error bars.~The BSSs in the wide-field catalog are the ones selected by \cite{kra10}, i.e. stars on the BSS region of all ($B$-$V$), ($V$-$I$) and ($B$-$I$) CMDs.~We then apply an additional magnitude cut ($I<19.45$ mag) in order to hold the same faint magnitude limit used with the inner sample (see Fig~\ref{cmd1})\footnote{The F814W$<$19.5 mag limit from Fig.~1 is checked using the ``$I$ $ground$" magnitude on the ACS catalog and is found to translate to $I < 19.45$ mag.}.~Considering these selection steps, combining both samples is hence acceptable for our purposes.~We find that the BSS fraction is maximal in the center and drops rapidly with radius reaching near to zero at $r\!\sim\!10\,r_c$.~A subsequent rising in the fraction profile is discernible, although this is only caused by the detection of two BSSs alone (see the inset panel) and therefore not strongly supported statistically.~The absence of a clear outer layer suggests that the majority of the BSS population has already been affected by dynamical friction. Hence, in the framework constructed by \cite{fer12}, NGC\,1261 would classify as a dynamically old cluster and would be grouped in late-{\sc Family\,II}/{\sc Family\,III} , not surprisingly similar to M\,30 and NGC\,362.~Moreover, the half-mass relaxation time is about $10^8$ years, which is actually shorter than that of NGC\,362, $\log t=8.7$ and M\,30, $\log t=9.2$, as found in \cite{pau10}. \begin{figure}[t!] \centering \includegraphics[width=9cm]{fig5.pdf} \caption{\small{Ratio of BSS to SGB stars as a function of radial distance.~The inset panel shows the number of BSSs and SGB stars. Both panels have a dashed vertical line indicating the approximate location of the cluster's tidal radius, according to \cite{pau10}}.} \label{BSS_rad_frac} \end{figure} However, the core structure of NGC 1261 does not show signatures of core collapse.~It is well approximated by a King model with a concentration parameter $c\!\approx\!1.2$ \citep{har96, pau10}.~It has a central luminosity density of 2.22 $L_{\odot}$pc$^{-3}$ \citep{pau10}, which is relatively low for GCs.~Moreover, the binary fraction radial profile found by \cite{mil12} shows a flat distribution, i.e.~without signs of mass segregation, contrary to the ones in M\,30 and NGC\,362, which are centrally peaked.~Therefore it is not surprising that the binary merger products, i.e.~the R-BSS population, in these GCs are also more centrally segregated than the R-BSS population in NGC\,1261.~This apparent evolutionary contradiction may not be as problematic as it first seems.~According to both Monte-Carlo dynamical models \citep[e.g.][]{heg08} and direct-integration N-body models \citep[e.g.][]{hur12}, clusters can go through core-collapse and then, if there is some energy source in the core, stay or pass through a long-lived post-core-collapse bounce state in which they do not show classic post core-collapse signatures.~The \citeauthor{heg08} models of M\,4 all went through core collapse at $t\!=\!8$\,Gyr, and then remained in a non-collapsed state for another $\sim\!2\!-\!3$ Gyr due to ``binary burning" in the core. \citeauthor{hur12} propose a binary black hole as an alternative central potential energy source.~We know today that black holes may be common in GCs \citep[e.g.][]{str12, cho13}, so that a binary black hole in the core of NGC\,1261 may be considered a valid possibility.\\ \section{Summary and Conclusions} We find that the inner BSS population in NGC\,1261 includes at least two distinct well defined sequences similar to what was found in M\,30 and NGC\,362, and as well includes a smaller group of BSSs that have unusually blue colours in the CMD, and which could be associated with a distinct coeval BSS population, if confirmed real.~The comparison with collisional stellar evolution models reveals that the B-BSS and eB-BSS sub-samples are consistent with a 2 Gyr and 0.2 Gyr old stellar collision-product population, respectively. This provides the grounds for considering NGC\,1261 an extremely valuable test laboratory for stellar collision and BSS formation models.~This observation along with evidence collected from the literature suggest as a preliminary interpretation for the dynamical history of NGC\,1261 that the cluster experienced a core-collapse phase about 2 Gyr ago, and that since then it has bounced through core oscillations.~The subsequent core oscillations occasionally created more BSSs during these short timescale processes when the central stellar density was particularly enhanced -- one such likely 0.2 Gyr ago. During these periods of core density enhancements the cluster likely burned some of its core binaries, thereby flattening their radial density distribution profile, and currently the cluster is likely in a post core-collapsed state which, according to different simulations, may appear as an unevolved GC. Follow-up spectroscopic characterization of the BSS sequences in NGC\,1261 is of utmost importance in order to confirm and better understand their origins and formation mechanisms, and in particular test the chemical abundance predictions related to different BSS formation models. \acknowledgments We thank the anonymous referee for comments that improved the presentation and quality of our results.~MS and THP gratefully acknowledge support from CONICYT through the ALMA-CONICYT Project No.~37070887, FONDECYT Regular Project No.~1121005, FONDAP Center for Astrophysics (15010003), and BASAL Center for Astrophysics and Associated Technologies (PFB-06), as well as support from {\it Deutscher Akademischer Austauschdienst} (DAAD). AS is supported by NSERC.~This research has made use of the Aladin plot tool and the TOPCAT table manipulation software, found at http://www.starlink.ac.uk/topcat/. {\it Facilities:} \facility{HST (ACS)}.
\section{Introduction} A sequence of integers $\{n_k\} \subset \mathbb{Z}$ is said to be \emph{universally $L^p$-good} if for every measure-preserving system $(X,\mu,T)$ and every $f \in L^p(X)$ the subsequence averages \[ A_N^{\{n_k\}} f := \frac{1}{N}\sum_{k=1}^N T^{n_k}f \] converge pointwise almost everywhere. In this language, Birkhoff's classical pointwise ergodic theorem \cite{0003.25602} states that the full sequence of integers is universally $L^1$-good. Obtaining pointwise convergence results for rougher, sparser sequences is much more challenging. For instance, Bourgain's Polynomial Ergodic Theorem \cite{MR1019960} states that the sequence $\{P(n)\}$, $P$ integer polynomial, is universally $L^p$-good for each $p>1$. Note that $\{P(n)\}$ are zero-Banach-density subsequences of the integers; in fact, Bourgain used a probabilistic method to find \emph{extremely} sparse universally good sequences. From now on $\{X_n\}$ will denote a sequences of independent $\{0,1\}$ valued random variables (on a probability space $\Omega$) with expectations $\sigma_n$. The \emph{counting function} $a_n(\omega)$ is the smallest integer subject to the constraint \[ X_1(\omega) + \dots + X_{a_n(\omega)}(\omega) = n.\] \begin{theorem}[{\cite[Proposition 8.2]{MR937581}}] Suppose \[ \sigma_n = \frac{ (\log \log n)^{B_p}}{n}, \ B_p > \frac{1}{p-1}, \ 1 < p \leq 2. \] Then, almost surely, $\{a_{n}\}$ is universally $L^p$-good. \end{theorem} In the years to follow random sequences became a widely used model for pointwise ergodic theorems. One indication at their amenability to analysis is LaVictoire's $L^{1}$ random ergodic theorem. \begin{theorem}[{\cite{MR2576702}}] Suppose $\sigma_n = n^{-a}$ with $0< a < 1/2$. Then, almost surely, $\{a_{n}\}$ is universally $L^1$-good. \end{theorem} Here, by the strong law of large numbers, almost surely \[ a_n(\omega)/ n^{\frac{1}{1-a}} \] converges to a non-zero number. For comparison, it is known that the sequences of $d$-th powers, $d>1$ integer, are universally $L^{1}$ bad \cite{MR2331324,MR2788358}. Random sequences have also been used as a model for multiple ergodic averages. Frantzikinakis, Lesigne, and Wierdl recently showed the following. \begin{theorem}[{\cite[Theorem 1.1]{MR3043589}}] \label{thm:flw} Suppose $\sigma_{n} = n^{-a}$, $0<a<1/14$. Then, almost surely, $(a_{n})_{n}$ has the following property: for every pair of measure preserving transformations $T,S$ on a probability space $X$ and any functions $f,g\in L^{\infty}(X)$ the averages \[ \sum_{n=1}^{N} g(S^{n}x) f(T^{a_{n}}x) \] converge pointwise almost everywhere. \end{theorem} It is noted in their paper that the linear sequence of powers $S^{n}$ can likely be replaced by other deterministic sequences, but their method of proof did not seem to allow this. In this article we prove a related result in which we are able to replace the linear sequence of powers by a sequence drawn from a more general class at the cost of weakening the result in several other respects. More precisely, with $0 < \epsilon< 1$ arbitrary but fixed, suppose $p : \mathbb{R} \to \mathbb{R}$ is a \emph{logarithmico-exponential} function which satisfies \begin{enumerate} \item the second-order difference relationship \[ p(x+y+z) - p(x+y) - p(x+z)+ p(x) = O(x^{\epsilon -1} yz) \] for $x,y, z >0$ (``big-O'' notation is recalled in the section on notation below); and \item for all $a(x) \in C \cdot \mathbb{Q} [x]$, the set of real constant multiples of rational polynomials, $\frac{|a(x) - p(x)|}{\log x} \to \infty$. \end{enumerate} Good examples of such functions are $p(x) = x^{1+\epsilon}$. We refer the reader to \cite{MR2145587} for a more complete discussion of \emph{logarithmico-exponential} functions; informally, these are all the functions one can get by combining real constants, the variable $x$, and the symbols $\exp, \ \log, \cdot,$ and $+$. (e.g.\ $x^{1/2} = \exp(1/2 \cdot \log x)$ and $x^\pi/\log \log x$ are both logarithmico-exponential.) Our main result is the following \begin{theorem}\label{MAIN} Suppose $\sigma_{n}=n^{-a}$, $0 < a < 1/2$, and $p$ is as above. Then, almost surely, the following holds: For each measure-preserving system $(X,\mu,T)$ and each $f \in L^1(X)$ the averages \[ \frac{1}{N} \sum_{n=1}^N e(p(n)) T^{a_n(\omega)}f \] converge to zero pointwise almost everywhere (here and later $e(t):= e^{2\pi i t}$). \end{theorem} Pointwise ergodic theorems with exponential polynomial weights are collectively known as Wiener--Wintner type theorems, see e.g.\ \cite{MR1995517} for linear polynomials and \cite{MR1257033} for general polynomials. If the random sequence $\{a_{n}\}$ is replaced by the linear sequence $\{n\}$ in Theorem~\ref{MAIN}, the result follows from the Wiener--Wintner theorem for Hardy field functions due to Eisner and the first author \cite{2014arXiv1407.4736E}. However, note that the full measure sets in our result depend on the choice of $p$. It would be interesting to remove this dependence. Also, the second order difference relation in the hypothesis of Theorem~\ref{MAIN} can likely be replaced by a polynomial growth assumption; this would require an inductive application of van der Corput's inequality. The structure of this paper is as follows:\\ In \textsection\ref{sec:preliminaries} we introduce a few preliminary tools, discuss our proof strategy, and reduce our theorem to proving Proposition \ref{key}; and \\ In \textsection\ref{sec:proof} we prove Proposition \ref{key}, thereby completing the proof of Theorem \ref{MAIN}. \subsection{Acknowledgments} We are grateful to Nikos Frantzikinakis, Rowan Killip, and Alexander Volberg for helpful conversations, and to Tanja Eisner for suggesting the collaboration. We thank the Hausdorff Research Institute for Mathematics for hospitality during the Trimester Program ``Harmonic Analysis and Partial Differential Equations''. \section{Preliminaries} \label{sec:preliminaries} \subsection{Notation and tools} With $X_n, \ \sigma_n$ as above, we let $Y_n := X_n - \sigma_n$. We will be dealing with sums of random variables, so we introduce the following compact notation: \[ S_{N} =\sum_{n=1}^{N} X_{N} \ \text{ and } \ S_{M,N}=\sum_{n=M}^{N} X_{n}.\] We also let \[ W_N := \sum_{n=1}^N \sigma_n,\] so that $W_N$ grows as $N^{1-a}$. We will make use of the modified Vinogradov notation. We use $X \lesssim Y$, or $Y \gtrsim X$ to denote the estimate $X \leq CY$ for an absolute constant $C$. If we need $C$ to depend on a parameter, we shall indicate this by subscripts, thus for instance $X \lesssim_\omega Y$ denotes the estimate $X \leq C_\omega Y$ for some $C_\omega$ depending on $\omega$. We also make use of big-O notation: we let $O(Y)$ denote a quantity that is $\lesssim Y$, and similarly $O_\omega(Y)$ a quantity that is $\lesssim_\omega Y$. The main probabilistic input in our argument is the following special case of Chernoff's inequality. \begin{lemma}[see e.g.\ \cite{MR2573797}] \label{lem:chernoff} Let $\{X_n\}$, $\{ \sigma_n \}$ be as above. There exists an absolute constant $c >0$ so that for each $A >0$, \[ \mathbb{P} \Big( | S_N - W_N | \geq A \Big) \lesssim \max\Big \{ \exp \Big(-c \frac{A^2}{W_N} \Big), \exp(-c A) \Big\}.\] Consequently, \[ \mathbb{P} \Big( |S_{N} - W_{N}| \geq \frac12 W_N \Big) \lesssim \exp(-c W_N) \lesssim \exp(-c N^{1-a}). \] \end{lemma} This also implies the following version of the law of large numbers: \begin{equation} \label{eq:lln} S_{N}/W_{N} \to 1 \quad\text{almost surely.} \end{equation} We will also need the Hilbert space van der Corput inequality. \begin{lemma}[{see e.g.\ \cite{MR3043589}}] Let $\{ v_{n}\} $ be a sequence in a Hilbert space $H$ and $1\leq M\leq N$. Then \begin{equation} \label{vdC} \Big\| \sum_{n=1}^{N} v_{n} \Big\|^{2}\ \leq 2 \frac{N}{M} \sum_{n=1}^{N} \| v_{n} \|^{2} + 4 \frac{N}{M} \sum_{m=1}^{M} \Big| \sum_{n=1}^{N-m} \< v_{n+m}, v_{n} \> \Big| \end{equation} \end{lemma} \subsection{Strategy} In proving his Random Ergodic Theorem, LaVictoire showed \cite{MR2576702} that on a set of full probability, $\Omega' \subset \Omega$, the maximal function \[ f \mapsto \sup_N \frac{1}{N} \sum_{n=1}^N T^{a_n(\omega)}|f|, \] is weakly bounded on $L^1(X)$. In particular, for $\omega \in \Omega'$ the set of $f \in L^1(X)$ for which the averages \[ \frac{1}{N} \sum_{n=1}^N e(p(n)) T^{a_n(\omega)} f \] tend to zero pointwise a.e.\ is closed in $L^1$. Hence it will be enough to prove pointwise convergence for $f \in L^\infty(X)$. Now, as observed in \cite{MR1325697}, for bounded functions it is enough to prove convergence along every lacunary sequence $\lac = (\lfloor \rho^k \rfloor, k\in\mathbb{N})$, where $\rho>1$ is taken from a countable sequence converging to $1$. We will fix some $\rho > 1$ throughout, and the averaging parameters $N$ are assumed to belong to $\lac$ unless mentioned otherwise. We follow a similar plan to \cite{MR3043589}. We will prove Theorem \ref{MAIN} by showing that almost surely, for every measure-preserving system $(X,\mu,T)$ and every $f\in L^{\infty}(X)$, the following chain of asymptotic equivalences holds $\mu$-almost everywhere: \begin{align} \frac{1}{N}\sum_{n=1}^N e(p(n)) T^{a_n}f &\approx\label{sim:def} \frac{1}{S_N}\sum_{n=1}^N X_n(\omega) e(p(S_n)) T^nf \\ &\approx\label{sim:lln} \frac{1}{W_N}\sum_{n=1}^N X_n(\omega) e(p(S_n)) T^nf \\ &\approx\label{sim:summable} \frac{1}{W_N}\sum_{n=1}^N \sigma_n e(p(S_n)) T^nf \\ &\approx\label{sim:ergodic-av} \bar{f} \cdot \frac{1}{W_N}\sum_{n=1}^N \sigma_n e(p(S_n)) \\ &\approx\label{sim:reverse} \bar{f} \cdot \frac{1}{N} \sum_{n=1}^N e(p(n)) \\ &\approx\label{sim:zero} 0. \end{align} Here, the symbol $\approx$ means that the difference converges to $0$ as $N\to\infty$ and $\bar{f} := \lim_N \frac{1}{N} \sum_{n=1}^{N} T^n f$ is the projection of $f$ onto the invariant factor of $T$. Let us now list the ingredients used to establish the above asymptotic equivalences. \begin{enumerate} \item[\eqref{sim:def}] holds because the right-hand side equals the left-hand side with $N$ replaced by $S_{N}$. \item[\eqref{sim:lln}] holds by \eqref{eq:lln}. \item[\eqref{sim:summable}] is the key to our argument. We isolate this crucial step in the following \begin{proposition}\label{key} In the setting of Theorem~\ref{MAIN}, almost surely the following holds: for each measure-preserving system $(X,\mu,T)$, and each $f \in L^2(X)$, the sequence \[ \Big\| \frac{1}{W_N} \sum_{n=1}^N Y_n(\omega) e(p(S_n)) T^nf \Big \|_{L^2(X)}^2 \] is summable over lacunary $N$, and in particular \[ \frac{1}{W_N} \sum_{n=1}^N Y_n(\omega) e(p(S_n)) T^nf \to 0 \quad \mu\text{-a.e.} \] \end{proposition} \item[\eqref{sim:ergodic-av}] for averages with weights $\sigma_{n}$ follows by the partial summation formula \[ \frac{1}{W_{N}} \sum_{n=1}^{N} \sigma_{n} a_{n} = \frac{N\sigma_{N}}{W_{N}} A_{N} + \sum_{M=1}^{N-1} \frac{M(\sigma_{M}-\sigma_{M+1})}{W_{N}} A_{M}, \quad A_{N} = \frac{1}{N} \sum_{n=1}^{N} a_{n}, \] from the following result on unweighted averages with $G=e\circ p$: \begin{lemma} \label{lem:ergodic-av} Suppose $0<a<1$. Then, almost surely, for every measure-preserving system $(X,\mu,T)$ and every $f\in L^{1}(X,\mu)$ pointwise $\mu$-a.e.\ we have \[ \frac{1}{N}\sum_{n=1}^N G(S_n) T^nf \approx \bar{f} \cdot \frac{1}{N}\sum_{n=1}^N G(S_n) \] for every bounded function $G:\mathbb{N}\to\mathbb{R}$ as $N\to\infty$. \end{lemma} This is a slight abstraction from \cite[Lemma 2.2]{MR3043589}, where a different function $G$ was specified (but its special form not used in the proof). For completeness, the proof is reproduced below. \item[\eqref{sim:reverse}] follows by applying the above steps in reverse order, with $f = 1_X$; and \item[\eqref{sim:zero}] reduces to a statement about trigonometric sums, namely $\frac{1}{N} \sum_{n=1}^N e(p(n)) \to 0$, which was proved in \cite[Theorem 1.3]{MR1269206}. \end{enumerate} \begin{proof}[Proof of Lemma~\ref{lem:ergodic-av}] By the usual maximal ergodic theorem, for each fixed $\omega$ the set of $f$ for which asymptotic equivalence holds a.e.\ is closed in $L^{1}(X)$. Since the equivalence is clear in the case $f=\bar f$ and in view of the splitting $L^{2}(X) = \{f=\bar f\} \oplus \overline{\{ Th-h, h\in L^{\infty}(X)\}}$, it suffices to consider the case when $f=Th-h$, $h\in L^{\infty}$, is a coboundary, so that in particular $\bar f = 0$. Since $f\in L^{\infty}$ in this case, it suffices to obtain equivalence for $N\in\lac$ with $\rho>1$ fixed but arbitrary. Summation by parts gives \[ \frac{1}{N}\sum_{n=1}^N G(S_n) T^n(h-Th) = O(\|G\|_{\infty}/N) + \frac{1}{N} \sum_{n=1}^{N-1} (G(S_n) - G(S_{n+1})) T^n h. \] The first summand is deterministic and converges to $0$. The second summand is $\mu$-a.e.\ bounded by \[ 2\|G\|_{\infty}\|h\|_{\infty} \frac{1}{N} \sum_{n=1}^{N-1} X_{n+1} \leq 2\|G\|_{\infty}\|h\|_{\infty} \frac{S_{N}}{N}, \] and this converges to $0$ almost surely in view of \eqref{eq:lln}. \end{proof} With this reduction complete, we now turn to the proof of Proposition \ref{key}. \section{Proof of Proposition \ref{key}} \label{sec:proof} Throughout this section, we will view $0< \delta \ll 1$ as a (small) floating parameter, whose precise value will be fixed at the end of the proof; $0 < \nu = \nu(\delta) = O(\delta)$ will be used to denote (possibly different) parameters (all of which grow linearly in $\delta$); $0< \kappa = O(\delta)$ will be used similarly. We begin with a criterion that guarantees that a bounded sequence $\{c_n\}$ is a good sequence of weights for a pointwise ergodic theorem along a lacunary sequence. \begin{lemma} \label{lem:crit} Let $0<a<b<1$ and fix $\rho>1$. Let $\{c_n\}$ be a bounded sequence such that the following holds: \begin{equation} \label{eq:c-sum} \sum_{n=1}^{N} |c_{n}| \lesssim N^{1-a}, \quad N\in\lac, \quad\text{and} \end{equation} \begin{equation} \label{eq:c-corr} \sum_{N\in\lac} N^{2a-1-b} \sum_{m=1}^{N^{b}} \Big| \sum_{n=N^{1-\delta}}^{N-m} c_{n+m} \bar c_{n} \Big| < \infty. \end{equation} Then for every measure-preserving system $(X,\mu,T)$ and $f\in L^{2}(X)$ we have \[ \sum_{N\in\lac} \Big\| \frac1{N^{1-a}} \sum_{n=1}^{N} c_{n} T^{n}f \Big\|_{L^2(X)}^{2} < \infty. \] \end{lemma} \begin{proof} Note that \eqref{eq:c-sum} with $N\in\lac$ implies \eqref{eq:c-sum} with $N\in\mathbb{N}$, and we obtain \begin{align*} \Big\| \frac1{N^{1-a}} \sum_{n=1}^{N^{1-\delta}} c_{n} T^{n}f \Big\|_{L^2(X)} &\leq \frac1{N^{1-a}} \sum_{n=1}^{N^{1-\delta}} |c_{n}| \| f \|_{L^2(X)}\\ &\lesssim \frac1{N^{1-a}} N^{(1-\delta)(1-a)} \| f \|_{L^2(X)} \\ &= N^{-\delta (1-a)} \| f \|_{L^2(X)}, \end{align*} so we may replace the sum in the conclusion of the lemma by $\sum_{n=N^{1-\delta}}^{N}$. Using van der Corput inequality \eqref{vdC} on the Hilbert space $H=L^{2}(X)$ with $M=N^{b}$, estimate \begin{multline} \label{eq:vdC1} \Big\| \frac1{N^{1-a}} \sum_{n=N^{1-\delta}}^{N} c_{n} T^{n}f \Big\|_{L^2(X)}^{2}\\ \lesssim N^{2a-2} \frac{N}{N^{b}} \sum_{n=N^{1-\delta}}^{N} \| c_{n} T^{n}f \|_{L^{2}(X)}^{2} + N^{2a-2} \frac{N}{N^{b}} \sum_{m=1}^{N^{b}} \Big| \sum_{n=N^{1-\delta}}^{N-m} \int_{X} c_{n+m} T^{n+m} f \bar c_{n} T^{n} \bar{f} \Big|. \end{multline} The first term in \eqref{eq:vdC1} is bounded by \[ N^{2a-2} \frac{N}{N^{b}} \sum_{n=N^{1-\delta}}^{N} | c_{n} |^{2} \|f\|_{L^2(X)}^{2}, \] and by the assumption \eqref{eq:c-sum} and boundedness of $(c_{n})$ this is $O(N^{a-b})$. By precomposing with $T^{-n}$, the second term in \eqref{eq:vdC1} is bounded by \[ N^{2a-1-b} \sum_{m=1}^{N^{b}} \Big| \sum_{n=N^{1-\delta}}^{N-m} c_{n+m} \bar c_{n} \Big| |\<T^{m}f,f\>_{L^2(X)}|, \] and this is summable by the assumption \eqref{eq:c-corr}. \end{proof} \begin{proposition} \label{prop:as-corr} Let $p:\mathbb{R}\to\mathbb{R}$ be a function such that \begin{equation} \label{eq:p} p(x+y+z) - p(x+y) = p(x+z) - p(x) + O(x^{\epsilon-1}yz) \end{equation} for $x,y,z>0$. Let also $0<a<1/2$ and fix $\rho>1$. Then there exists $b\in (a,1/2)$ such that, almost surely, the sequence $c_{n}=Y_{n} e(p(S_{n}))$ satisfies \eqref{eq:c-corr}. \end{proposition} \begin{proof} By Fubini's theorem it suffices to show that the expectation \[ N^{2a-1-b} \sum_{m=1}^{N^{b}} \underbrace{\mathbb{E} \Big| \sum_{n=N^{1-\delta}}^{N-m} Y_{n+m} e(p(S_{n+m})) Y_{n} e(-p(S_{n})) \Big|}_{=: I(m)} \] is summable along the lacunary sequence $N\in\lac$. By Cauchy--Schwarz we have \begin{equation} \label{eq:pre-vdC2} I(m)^{2} \leq \mathbb{E} \Big| \sum_{n=N^{1-\delta}}^{N-m} Y_{n} Y_{n+m} e(p(S_{n+m})-p(S_{n})) \Big|^{2}. \end{equation} Using the van der Corput inequality \eqref{vdC} with values in the Hilbert space $H=L^{2}(\Omega)$ and $R=N^{c}$, $0<c<1$ to be chosen later, we obtain the estimate \begin{align*} &I(m)^2 \leq I_1(m)^2 + I_2(m)^2 + I_3(m)^2 :=\\ &\quad \frac{N-m}{R} \sum_{n=N^{1-\delta}}^{N-m} \| Y_{n} Y_{n+m} e(p(S_{n+m})-p(S_{n})) \|_{L^{2}(\Omega)}^{2}\\ & + \frac{N-m}{R} \Big| \mathbb{E} \sum_{n=N^{1-\delta}}^{N-2m} Y_{n+2m} Y_{n+m} Y_{n+m} Y_{n} e(p(S_{n+2m})-p(S_{n+m})-p(S_{n+m})+p(S_{n})) \Big| \\ & + \frac{N-m}{R} \sum_{r=1, r \neq m}^{R} \Big| \mathbb{E} \sum_{n=N^{1-\delta}}^{N-m-r} Y_{n+r} Y_{n+r+m} Y_{n} Y_{n+m} e(p(S_{n+r+m})-p(S_{n+r})-p(S_{n+m})+p(S_{n})) \Big|. \end{align*} The task is now to show that, uniformly in $m \leq N^b$, we have \[ I_j(m)^2 \lesssim N^{2 - 4a - \kappa} \text{ for each } j = 1,2,3, \] for some $\kappa = \kappa(\delta,a,b,c) > 0$. To this end we estimate the first term, $I_1(m)^2$, by \[ \frac{N}{R} \sum_{n=N^{1-\delta}}^{N-m} \| Y_{n} \|_{L^{2}(\Omega)}^{2} \| Y_{n+m} \|_{L^{2}(\Omega)}^{2} \] by independence; this is bounded by \[ \frac{N}{R} \sum_{n=N^{1-\delta}}^{N-m} \sigma_{n} \sigma_{n+m} \lesssim N^{1-c} N^{1-2a} < N^{2 - 4a - \kappa} \] provided we take $ 2 a < c < 1$. We next turn to $I_2(m)^2$, which contributes at most \[ \frac{N}{R} \mathbb{E} \sum_{n=N^{1-\delta}}^{N-2m} | Y_{n+m} Y_{n+2m} Y_{n} Y_{n+m} |, \] and by independence this is bounded by \begin{align*} \frac{N}{R} \sum_{n=N^{1-\delta}}^{N-2m} \mathbb{E} |Y_{n+m}|^{2} \cdot \mathbb{E} |Y_{n+2m}| \cdot \mathbb{E} |Y_{n}| &\lesssim N^{1-c} \sum_{n=N^{1-\delta}}^{N-2m} \sigma_{n} \sigma_{n+m} \sigma_{n+2m} \\ &\lesssim N^{1-c} N^{1-3a+\nu} \\ &\lesssim N^{2 - 4a - \kappa}, \end{align*} provided $c > 2a$ (from above) and $\nu = \nu(\delta) > 0$ is taken sufficiently small. ($\nu$ arises from the possibility that $3a > 1$, in which case we may take e.g.\ $\nu = (3a -1) \delta$.) The contribution of this term is also acceptable. It remains to estimate $I_3(m)^2$, which we write in the form \[ I_3(m)^2 = \frac{N-m}{R} \sum_{r=1, r \neq m}^{R} \Big| \mathbb{E} \sum_{n=N^{1-\delta}}^{N-m-r} Y_{n+r} Y_{n+r+m} Y_{n} Y_{n+m} e(p(S_{n+s+t})-p(S_{n+t})-p(S_{n+s})+p(S_{n})) \Big|, \] with $s=\min(r,m)$ and $t=\max(r,m)$. To recover independence we apply \eqref{eq:p} with \[ x=S_{n+t-1}, \ y=X_{n+t}, \ z=S_{n+t+1,n+t+s}, \] to the first two summands in the argument of $e$. This gives the estimate \begin{align*} &\frac{N-m}{R} \sum_{r=1, r \neq m}^{R} \Big| \mathbb{E} \sum_{n=N^{1-\delta}}^{N-m-r} Y_{n+r} Y_{n+r+m} Y_{n} Y_{n+m} e(p(S_{n+t-1}+S_{n+t+1,n+t+s})-p(S_{n+t-1})-p(S_{n+s})+p(S_{n})) \Big|\\ &+ \frac{N-m}{R} \sum_{r=1, r \neq m}^{R} \mathbb{E} \sum_{n=N^{1-\delta}}^{N-m-r} \Big| Y_{n+r} Y_{n+r+m} Y_{n} Y_{n+m} \min(O(S_{n+t-1}^{\epsilon-1} X_{n+t} S_{n+t+1,n+t+s}),1) \Big|. \end{align*} The main feature of this splitting is that the exponential in the first term does not depend on $X_{n+t}$, so $Y_{n+t}$ is independent from all other terms. Therefore the first term vanishes identically. The second term is estimated by \begin{align*} &\frac{N}{R} \sum_{r=1, r \neq m}^{R} \sum_{n=N^{1-\delta}}^{N-m-r} \mathbb{E} \Big( | Y_{n+r} Y_{n+r+m} Y_{n} Y_{n+m} | \cdot \min(S_{n+t-1}^{\epsilon-1} S_{n+t+1,n+t+s},1) \Big)\\ &\leq \frac{N}{R} \sum_{r=1, r \neq m}^{R} \sum_{n=N^{1-\delta}}^{N-m-r} \mathbb{E} \Big( | Y_{n+r} Y_{n+r+m} Y_{n} Y_{n+m} | \cdot \min(S_{n+t-1}^{\epsilon-1} (S_{n+t+1,n+t+s-1}+1),1) \Big) \end{align*} By Lemma~\ref{lem:chernoff} this is bounded by \[ \frac{N}{R} \sum_{r=1, r \neq m}^{R} \sum_{n=N^{1-\delta}}^{N-m-r} \mathbb{E} \Big( | Y_{n+r} Y_{n+r+m} Y_{n} Y_{n+m} | \cdot \min((W_{n+t-1})^{\epsilon-1} (S_{n+t+1,n+t+s-1}+1), 1_{E_{n+t-1}}) \Big), \] where $E_{n}$ is an exceptional set of measure $\lesssim \exp(-c n^{1-a})$. At this point we estimate the minimum by a sum. We consider first the non-exceptional part. All remaining random variables are independent, so we get the estimate \begin{align*} &\frac{N}{R} \sum_{r=1, r \neq m}^{R} \sum_{n=N^{1-\delta}}^{N-m-r} \sigma_{n+r} \sigma_{n+r+m} \sigma_{n} \sigma_{n+m} (n+t-1)^{(1-a)(\epsilon-1)} \Big( \sum_{j=n+t+1}^{n+t+s-1}\sigma_{j}+1 \Big)\\ & \qquad \leq \frac{N}{R} \sum_{r=1, r \neq m}^{R} \sum_{n=N^{1-\delta}}^{N-m-r} n^{-4a} n^{(1-a)(\epsilon-1)} (n^{-a} N^{b}+1)\\ &\lesssim \frac{N}{R} \sum_{r=1, r \neq m}^{R} N^{b} \sum_{n=N^{1-\delta}}^{N-m-r} n^{-4a} n^{(1-a)(\epsilon-1)} n^{-a}\\ &\lesssim N^{2 - 4a + ( b- a + \nu) + (1-a)(\epsilon -1)} \\ &\lesssim N^{2 - 4a - \kappa}, \end{align*} provided that $b$ is taken sufficiently close to $a$ and $\delta$ is sufficiently small, since $(1-a)(\epsilon-1)<0$. Finally, for the exceptional part we have superpolynomial decay in $N$. \end{proof} Thus we have verified that the assumption \eqref{eq:c-corr} of Lemma~\ref{lem:crit} holds almost surely in the setting of Proposition~\ref{key}. The missing assumption \eqref{eq:c-sum} also holds almost surely because \[ \sum_{n=1}^{N} |Y_{n}| \leq S_{N} + W_{N} \] and in view of \eqref{eq:lln}. This completes the proof of Proposition~\ref{key} and hence of Theorem~\ref{MAIN}. \printbibliography \end{document}
\section{Introduction} In the present paper we prove existence results for solutions to nonlinear elliptic Neumann problems whose prototype is \begin{equation} \label{pb0} \begin{cases} -\Delta_{p} u -{\operatorname{div}} (c(x)|u|^{p-2}u) =f & \text{in}\ \Omega, \\[.1cm] \left( |\nabla u|^{p-2}\nabla u+ c(x)|u|^{p-2}u \right)\cdot\underline n=0 & \text{on}\ \partial \Omega ,% \end{cases}% \end{equation}% where $\Omega$ is a bounded domain of $\mathbb R^{N}$, $N\geq 2$, with Lipschitz boundary, $1< p\le N$ , $\underline n$ is the outer unit normal to $\partial \Omega$, the datum $f$ belongs to $L^{1}(\Omega)$ and satisfies the compatibility condition $\int_\Omega f =0$. Finally the coefficient $c(x)$ belongs to an appropriate Lebesgue space. When $c(x)=0$ and $f$ is an element of the dual space of the Sobolev space $W^{1,p}(\Omega)$, the existence and uniqueness (up to additive constants) of weak solutions to problem \eqref{pb0} is consequence of the classical theory of pseudo monotone operators (cfr. \cite{LL}, \cite{Lions}). But if $f$ is just an $L^1-$function, and not more an element of the dual space of $W^{1,p}(\Omega)$, one has to give a meaning to the notion of solution. \noindent When Dirichlet boundary conditions are prescribed, various definitions of solution to nonlinear elliptic equations with right-hand side in $L^1$ or measure have been introduced. In \cite{BBGGPV}, \cite{Aglio}, \cite{LM}, \cite{murat94} different notions of solution are defined even if they turn out to be equivalent, at least when the datum is an $L^{1}-$ function. The study of existence or uniqueness for Dirichlet boundary value problems has been the object of several papers. We just recall that the linear case has been studied in \cite{St}, while the nonlinear case began to be faced in \cite{BG1} and \cite{BG2} and was continued in various contributions, including \cite{AM1}, \cite{BeGu}, \cite{BBGGPV}, \cite{BMMP1}, \cite{BMMP}, \cite{DMOP}, \cite{Aglio}, \cite{GM2}, \cite{GM1}; mixed boundary value problems have been also studied (see \cite{BeGu}). In the present paper we refer to the so-called renormalized solutions (see \cite{DMOP}, \cite{LM}, \cite{murat94}) whose precise definition is recalled in Section 2. The existence for Neumann boundary value problems with $L^1-$ data when $c=0$ has been treated in various contests. In \cite{AMST97}, \cite{Ciabr}, \cite{Droniou00}, \cite{DV} and \cite{Prignet97} the existence of a distributional solution which belongs to a suitable Sobolev space and which has null mean value is proved. Nevertheless when $p$ is close to 1, i.e. $p\le 2-1/N$, the distributional solution to problem \eqref{pb0} does not belong to a Sobolev space and in general is not a summable function; this implies that its mean value has not meaning. This difficulty is overcome in \cite{Rako} by considering solutions $u$ which are not in $L^1(\Omega)$, but for which $\Phi(u)$ is in $L^1(\Omega)$, where $\Phi(t)=\int_{0}^{t}\frac{ds}{(1+|s|)^\alpha}$ with appropriate $\alpha>1$. In \cite{ACMM} the case where both the datum $f$ and the domain $\Omega$ are not regular is studied and solutions whose median is equal to zero are obtained with a natural process of approximations. We recall that the median of $u$ is defined by \begin{equation}\label{-520bis} {\rm med } (u) = \sup \{t\in \mathbb R : \operatorname{meas}\{u>t\} \geq \operatorname{meas}(\Omega)/2\}\, . \end{equation} Neumann problems have been studied by a different point of view in \cite{FM2, FM}. In this paper we face two difficulties: one due to the presence of the lower order term $-{\operatorname{div}} (c(x)|u|^{p-2}u))$ and the other due to the low integrability properties of the datum $f$. Our main result is Theorem \ref{exist_renorm} which asserts the existence of a renormalized solution to \eqref{pb0} having ${\operatorname{med}}(u)=0$. Its proof, contained in Section 4, is based on an usual procedure of approximation which consists by considering problems of type \eqref{pb0} having smooth data which strongly converge to $f$ in $L^1$. For such a sequence of problems we prove in Section 3 an existence results for weak solutions which is obtained by using a fixed point arguments. A priori estimates allow to prove that these weak solutions converge in some sense to a function $u$ and a delicate procedure of passage to the limit allows to prove that $u$ is a renormalized solution to \eqref{pb}. In Section 5 we give a stability result and we prove that, under larger assumptions on the summability of $f$, a renormalized solution to \eqref{pb} is in turn a weak solution to the same problem. At last Section 6 is concerned with Neumann problems with a zero order term; adapting the proof of Theorem \ref{exist_renorm} allows to derive an existence result for this type of operators. \section{Assumptions and definitions} Let us consider the following nonlinear elliptic Neumann problem \begin{equation} \label{pb} \left\{ \begin{array}{lll} -\mbox{div}\left( \operatorname{\mathbf{a}}\left( x, u, \nabla u\right)+ \Phi (x,u) \right) =f & & \text{in}\ \Omega, \\ \left( \operatorname{\mathbf{a}}\left( x, u, \nabla u\right)+ \Phi (x,u) \right)\cdot\underline n=0& & \text{on}\ \partial \Omega ,% \end{array}% \right. \end{equation}% where $\Omega $ is a connected open subset of $\mathbb{R}^{N}$, $N\ge 2$, having finite Lebesgue measure and Lipschitz boundary, $\underline n$ is the outer unit normal to $\partial \Omega$. We assume that $p$ is a real number such that $1<p \leq N$ and \[ \operatorname{\mathbf{a}} :\Omega \times\mathbb{R}\times \mathbb{R}^{N}\rightarrow \mathbb{R}^{N}\, , \] \[ \Phi:\Omega \times \mathbb{R}\rightarrow \mathbb{R}^N \] are Carath\'{e}odory functions. Moreover $\operatorname{\mathbf{a}}$ satisfies: \begin{equation} \operatorname{\mathbf{a}}\left( x, s, \xi \right) \cdot \xi \geq \alpha \left\vert \xi \right\vert ^{p},\quad \forall s\in\mathbb R,\ \forall \xi\in\mathbb R^{N},\ \text{a.e. in $\Omega$}\label{ell} \end{equation} where $\alpha>0$ is a given real number; \begin{equation} \left( \operatorname{\mathbf{a}}\left( x, s, \xi \right) -\operatorname{\mathbf{a}}\left( x,s, \eta \right) \right) \cdot \left( \xi -\eta \right) \geq 0 \label{mon} \end{equation} $\forall s\in\mathbb R$, $\forall \xi,\eta\in\mathbb R^{N}$ with $\xi\neq \eta$ and a.e. in $\Omega$; \par \noindent for any $k>0$ there exist $a_{k}>0$ and $b_k $ belonging to $L^{p'}(\Omega)$ such that \begin{equation} \label{growth} \left\vert \operatorname{\mathbf{a}}\left( x,s, \xi \right) \right\vert \leq a_{k}\left\vert \xi \right\vert ^{p-1}+b_k(x),\quad \forall |s|<k,\ \forall \xi\in\mathbb R^{N},\ \text{a.e. in $\Omega$.} \end{equation} \par We assume that $\Phi$ satisfies the following growth condition \begin{equation} | \Phi( x,s)|\le c(x) ( 1+|s|^{p-1}) \label{growthphi} \end{equation} $\forall s \in \mathbb{R}$, a.e. in $\Omega$, with $c\in L^{\frac{N}{p-1}}(\Omega)$ if $p<N$ and $c\in L^{q}(\Omega)$ with $q>N/(N-1)$ if $p=N$. \par Finally we assume that the datum $f$ is a measurable function in a Lebesgue space $L^r(\Omega)$, $1\le r\le +\infty$, which belongs to the dual space of the classical Sobolev space $W^{1,p}(\Omega)$ or is just an $L^1-$ function. Moreover it satisfies the compatibility condition \begin{equation} \int_{\Omega}f\,dx=0.\label{comp} \end{equation} \par As explained in the Introduction we deal with solutions whose median is equal to zero. Let us recall that if $u$ is a measurable function, we denote the median of $u$ by \begin{equation} {\operatorname{med}} (u) = \sup \left\{ t\in \mathbb R:\operatorname{meas} \{x\in \Omega : u(x)>t\} >\frac{\operatorname{meas}(\Omega)}{2} \right\}. \end{equation} Let us explicitely observe that if ${\operatorname{med}}(u)=0$ then \begin{gather*} \operatorname{meas} \{x\in \Omega : u(x)>0\} \le\frac{\operatorname{meas}(\Omega)}{2}, \\ \operatorname{meas} \{x\in \Omega : u(x)<0\} \le\frac{\operatorname{meas}(\Omega)}{2} \,. \end{gather*} In this case a Poincar\'e-Wirtinger inequality holds (see e.g. \cite{Z}): \begin{proposition} If $u\in W^{1,p}(\Omega) $, then \begin{equation} \|u-{\operatorname{med}} (u)\|_{L^p(\Omega)}\le C\|\nabla u \|_{(L^p(\Omega))^N} \label{poincare} \end{equation} where $C$ is a constant depending on $p$, $N$, $\Omega$. \end{proposition} As pointed out in the Introduction, when the datum $f$ is not an element of the dual space of the classical Sobolev space $W^{1,p}(\Omega)$ or is just an $L^1$-function, the classical notion of weak solution does not fit. We will refer to the notion of renormalized solution to \eqref{pb} (see \cite{DMOP,murat94} for elliptic equations with Dirichlet boundary conditions) which we give below. In the whole paper, $T_{k}$, $k\ge 0$, denotes the truncation at height $ k$ that is $T_{k}(s)=\min(k,\max(s,-k))$, $\forall s\in\mathbb R$. \begin{definition}\label{defrenorm} A real function $u$ defined in $\Omega$ is a renormalized solution to \eqref{pb} if \begin{gather} \label{def1} \text{ $u$ is measurable and finite almost everywhere in $\Omega$,} \\ \label{def2} T_{k}(u)\in W^{1,p}(\Omega), \text{ for any $k>0$,} \\ \label{def3} \lim_{n\rightarrow +\infty }\frac{1}{n} \int_{\{ x\in\Omega;\, |u(x)|<n\}} \operatorname{\mathbf{a}}(x,u,\nabla u) \nabla u \,dx = 0 \end{gather} and if for every function $h$ belonging to $W^{1,\infty}(\mathbb R)$ with compact support and for every $\varphi\in L^{\infty}(\Omega)\cap W^{1,p}(\Omega)$ we have \begin{multline} \int_{\Omega} h(u) \operatorname{\mathbf{a}} (x,u, \nabla u) \nabla \varphi dx + \int_{\Omega} h'(u) \operatorname{\mathbf{a}} (x,u, \nabla u) \nabla u \varphi dx \label{def4}\\ + \int_{\Omega} h(u)\Phi (x, u) \nabla \varphi dx + \int_{\Omega} h'(u)\Phi (x, u) \nabla u \varphi dx \int_\Omega f\varphi h(u) dx. \end{multline} \end{definition} \begin{remark} A renormalized solution is not an $L^1_{loc}(\Omega)$-function and therefore it has not a distributional gradient. Condition \eqref{def2} allows to define a generalized gradient of $u$ according to Lemma 2.1 of \cite{BBGGPV}, which asserts the existence of a unique measurable function $v$ defined in $\Omega$ such that $\nabla T_k(u)=\chi_{\{|u|<k \}}v$ a.e. in $\Omega$, $\forall k>0$. This function $v$ is the generalized gradient of $u$ and it is denoted by $\nabla u$. Equality \eqref{def4} is formally obtained by using in \eqref{pb} the test function $\varphi h(u)$ and by taking into account Neumann boundary conditions. Actually in a standard way one can check that every term in \eqref{def4} is well-defined under the structural assumptions on the elliptic operator. \end{remark} \begin{remark} It is worth noting that growth assumption \eqref{growthphi} on $\Phi$ together with \eqref{def1}--\eqref{def3} allow to prove that any renormalized solution $u$ verifies \begin{equation} \label{ermk1} \lim_{n\rightarrow +\infty} \frac{1}{n} \int_{\Omega} |\Phi(x,u)|\times |\nabla T_{n}(u)| dx = 0. \end{equation} Without loss of generality we can assume that ${\operatorname{med}}(u)=0$. Growth assumption \eqref{growthphi} implies that \begin{equation*} \int_{\Omega} |\Phi (x, u)| \times |\nabla T_{n}(u)| dx \leq \frac{1}{n} \int_{\Omega} c(x) (1+|T_{n}(u)|)^{p-1} |\nabla T_{n}(u)| dx. \end{equation*} In the case $N>p$, using H\"older inequality we obtain \begin{equation}\label{star} \begin{split} \int_{\Omega} c(x) & (1+|T_{n}(u)|)^{p-1} |\nabla T_{n}(u)| dx \\ & \leq C \| c\|_{L^{N/(p-1)}(\Omega)} (1+ \| T_{n}(u)\|^{p-1}_{L^{p^{*}}(\Omega)}) \| \nabla T_{n}(u)\|_{(L^{p}(\Omega))^{N}}. \end{split} \end{equation} Since ${{\operatorname{med}}(T_{n}(u))}=0$, by Poincar\'e--Wirtinger inequality, i.e. Proposition 2.1, and Sobolev embedding theorem it follows that \begin{equation*} \begin{split} \int_{\Omega} c(x) & (1+|T_{n}(u)|)^{p-1} |\nabla T_{n}(u)| dx \\ & \leq C \| c\|_{L^{N/(p-1)}(\Omega)} (1+ \| \nabla T_{n}(u)\|^{p-1}_{(L^{p}(\Omega))^{N}}) \| \nabla T_{n}(u)\|_{(L^{p}(\Omega))^{N}} \end{split} \end{equation*} where $C>0$ is a generic constant independent of $n$. Therefore Young inequality leads to \begin{equation}\label{ermk2} \begin{split} \frac{1}{n} \int_{\Omega} c(x) & (1+|T_{n}(u)|)^{p-1} |\nabla T_{n}(u)| dx \\ & \leq \frac{C}{n} \| c\|_{L^{N/(p-1)}(\Omega)} (1+ \| \nabla T_{n}(u)\|^{p}_{(L^{p}(\Omega))^{N}} ) \end{split} \end{equation} In the case $N=p$ a similar inequality involving $ \| c\|_{L^{q}(\Omega)}$ with $q>N/(N-1)$ occurs. Due to the coercivity of the operator $\operatorname{\mathbf{a}}$ and to \eqref{def3} we have \begin{equation*} \lim_{n\rightarrow+\infty}\frac{1}{n} \int_{\Omega} |\nabla T_{n}(u)|^{p} dx =0. \end{equation*} By \eqref{star} and \eqref{ermk2} we conclude that \eqref{ermk1} holds. \end{remark} \section{A basic existence result for weak solutions} In this section we assume more restrictive conditions on the right-hand side $f$, on $\Phi$ and on the operator $\operatorname{\mathbf{a}}$ in order to prove the existence of a weak solution $u$ to problem \eqref{pb}, that is \begin{gather*} u \in W^{1,p}(\Omega), \\ \int_{\Omega} \operatorname{\mathbf{a}}(x,u,\nabla u)\nabla v dx +\int_{\Omega} \Phi(x,u) \nabla v dx =\int_{\Omega} fv dx \end{gather*} for any $v\in W^{1,p}(\Omega)$. We assume \begin{gather} \label{3eq1} f\in L^{r}(\Omega)\cap (W^{1,p}(\Omega))' \\ \label{3eq2} |\Phi(x,s)| \leq c(x) \quad \text{$\forall s\in\mathbb R$, a.e. in $\Omega$} \end{gather} with $c\in L^{\infty}(\Omega)$. Moreover the operator $\operatorname{\mathbf{a}}$ satisfies \begin{equation} \label{3eq3a} \left( \operatorname{\mathbf{a}}\left( x, s, \xi \right) -\operatorname{\mathbf{a}}\left( x,s, \eta \right) \right) \cdot \left( \xi -\eta \right) >0 \end{equation} $\forall s\in\mathbb R$, $\forall \xi,\eta\in\mathbb R^{N}$ with $\xi\neq \eta$ and a.e. in $\Omega$; \begin{equation} \label{3eq3} |\operatorname{\mathbf{a}}(x,s,\xi)|\leq a_{0}(|\xi|^{p-1}+|s|^{p-1}) +a_{1}(x) \quad \forall s\in\mathbb R,\, \forall \xi\in \mathbb R^{N},\ \text{a.e. in $\Omega$}, \end{equation} with $a_{0}>0$, $a_{1}\in L^{p'}(\Omega)$. \begin{theorem} \label{exist_weak} Assume that \eqref{ell}, \eqref{3eq1}--\eqref{3eq3} and \eqref{comp} hold. There exists at least one weak solution $u$ to problem \eqref{pb} having ${\operatorname{med}}(u)=0$. \end{theorem} \begin{proof} The proof relies on a fixed point argument. \par Let $v\in L^p(\Omega)$. Due to \eqref{ell}, \eqref{3eq3a} and \eqref{3eq3}, $(x,\xi)\in \Omega\times \mathbb R^{N}\mapsto \operatorname{\mathbf{a}}(x,v(x),\xi)$ is a strictly monotone operator and verifies \[ |\operatorname{\mathbf{a}}(x,v(x),\xi)|\leq a_{0}(|\xi|^{p-1}+|v(x)|^{p-1}) +a_{1}(x) \quad \forall \xi\in \mathbb R^{N},\ \text{a.e. in $\Omega$}. \] Since $\Phi(x,v(x))\in (L^{\infty}(\Omega))^{N}$, classical arguments (see e.g. \cite{LL}, \cite{Lions}) allow to deduce that there exists a unique $ u$ such that \begin{equation} \label{3eq4} u \in W^{1,p}(\Om), \qquad {\operatorname{med}}(u)=0 \end{equation} and \begin{equation} \int_{\Omega} \operatorname{\mathbf{a}}(x,v,\nabla u)\nabla \varphi \, dx =\int_{\Omega}f\varphi \, dx - \int_{\Omega} \Phi(x,v)\nabla \varphi \, dx\, , \quad \forall \varphi \in W^{1,p}(\Omega)\,. \label{weak_v} \end{equation} It follows that we can consider the functional $\Gamma : L^p(\Omega)\longrightarrow L^p(\Omega) $ defined by \[ \Gamma (v)=u\, , \qquad \forall v\in L^p(\Omega), \] where $u$ is the unique element of $W^{1,p}(\Omega)$ verifying \eqref{3eq4} and \eqref{weak_v}. We now prove that $\Gamma$ is a continuous and compact operator. \par Let us begin by proving that $\Gamma$ is continuous. Let $v_n\in L^p(\Omega)$ such that $v_n\rightarrow v$ in $L^p(\Omega)$. Up to a subsequence (still denoted by $v_n$) $v_n \rightarrow v$ a.e. in $\Omega$. Let $u_n=\Gamma (v_n) $ belonging to $W^{1,p}(\Omega)$ such that ${\operatorname{med}}(u_n)=0$ and such that \eqref{weak_v} holds with $v_{n}$ in place of $v$. \noindent Choosing $\varphi =u_n$ as test function in \eqref{weak_v} and using \eqref{ell} we obtain that \[ \alpha \int_{\Omega} |\nabla u_{n}|^{p} dx \leq \int_{\Omega} | fu_{n}| dx + \int_{\Omega} |\Phi(x,v_{n}) \nabla u_{n}| dx . \] Since ${\operatorname{med}}(u_n)=0$, from Poincar\'e-Wirtinger inequality \eqref{poincare}, \eqref{3eq1} and \eqref{3eq2} Young inequality and Sobolev embedding theorem lead to \begin{equation} \label{3eq5} \int_{\Omega} |\nabla u_{n}|^{p} dx \leq M \end{equation} where $M>0$ is a constant independent of $n$. Using again \eqref{poincare}, it follows that $u_{n}$ is bounded in $W^{1,p}(\Omega)$. As a consequence and in view of \eqref{3eq3}, there exists a subsequence (still denoted by $u_n$), a measurable function $u$ and a field $\sigma$ belonging to $(L^{p'}(\Omega))^{N}$ such that \begin{gather} \label{conv1} u_n\rightharpoonup u\quad\text{\rm weakly in } W^{1,p}(\Om),\\ \label{conv3} u_n\to u\quad\text{ strongly in }L^{p}(\Omega), \\ \label{conv2} u_n\to u\quad\text{\rm a.e. in }\Omega,\\ \label{conv3a} \operatorname{\mathbf{a}}(x,v_{n},\nabla u_{n})\rightharpoonup \sigma\quad\text{ weakly in } (L^{p'}(\Omega))^{N}. \end{gather} Since ${\operatorname{med}}(u_{n})=0$ for any $n$ and since $u\in W^{1,p}(\Omega)$ the point-wise convergence of $u_{n}$ to $u$ implies that $\text{med}(u)=0$. To get the continuity of $\Gamma$ it remains to prove that $u=\Gamma (v)$ that is $u$ satisfies \eqref{weak_v}. Using \eqref{weak_v} with $v_{n}$ in place of $v$ and the test function $u_{n}-u$ we have \begin{multline}\label{new} \int_{\Omega} \operatorname{\mathbf{a}}(x,v_n,\nabla u_n)(\nabla u_n -\nabla u) dx = \int_{\Omega} f (u_{n}-u) dx \\ - \int_{\Omega} \Phi(x,v_{n})(\nabla u_{n}-\nabla u) dx. \end{multline} The point-wise convergence of $v_{n}$ and assumption \eqref{3eq2} imply that $\Phi(x,v_{n})$ converges to $\Phi(x,v)$ almost everywhere in $\Omega$ and in $L^{\infty}$ weak-* as $n$ goes to infinity. Therefore from \eqref{conv1} and \eqref{conv3}, passing to the limit in the right-hand side of \eqref{new}, we obtain \begin{equation}\label{minty0} \lim_{n\rightarrow +\infty} \int_{\Omega}\operatorname{\mathbf{a}}(x,v_n,\nabla u_n)(\nabla u_n -\nabla u) dx = 0. \end{equation} Let us recall the classical arguments, so-called Minty arguments, (see \cite{LL}, \cite{Lions}) which allow to identify $\sigma$ with $\operatorname{\mathbf{a}}(x,v,\nabla u)$. Let $\phi$ belonging to $(L^{\infty}(\Omega))^{N}$. Due to assumption \eqref{3eq3} and the convergence of $v_{n}$ the Lebesgue theorem shows that for any $t\in\mathbb R$ \begin{equation*} \operatorname{\mathbf{a}}(x,v_{n},\nabla u+t\phi) \rightarrow \operatorname{\mathbf{a}}(x,v,\nabla u+t\phi) \text{ strongly in $(L^{p'}(\Omega))^{N}$}. \end{equation*} By \eqref{conv3a} and \eqref{minty0}, it follows that for any $t\in\mathbb R$ \begin{align*} \lim_{n\rightarrow +\infty} \int_{\Omega} [\operatorname{\mathbf{a}}(x,v_n,\nabla u_n)- & \operatorname{\mathbf{a}}(x,v_{n},\nabla u + t\phi)](\nabla u_n -\nabla u-t\phi ) dx \\ & = \int_{\Omega} [\sigma - \operatorname{\mathbf{a}}(x,v,\nabla u+t\phi)] t\phi dx. \end{align*} Using the monotone character \eqref{3eq3a} of $\operatorname{\mathbf{a}}$ we obtain that for any $t\neq 0$ \begin{equation*} {\rm sign}\,(t)\int_{\Omega} [\sigma - \operatorname{\mathbf{a}}(x,v,\nabla u+t\phi)] \phi dx \geq 0. \end{equation*} Since $\operatorname{\mathbf{a}}(x,v,\nabla u +t \phi)$ converges strongly to $\operatorname{\mathbf{a}}(x,v,\nabla u)$ in $(L^{p'}(\Omega))^{N}$ as $t$ goes to zero, letting $t\rightarrow 0$ in the above inequality leads to \begin{equation*} \int_{\Omega} [\sigma-\operatorname{\mathbf{a}}(x,v,\nabla u)] \phi dx =0 \end{equation*} for any $\phi$ belonging to $(L^{\infty}(\Omega))^{N}$. We easily conclude that \begin{equation} \label{conv4} \sigma= \operatorname{\mathbf{a}}(x,v,\nabla u). \end{equation} By using \eqref{conv3a} and \eqref{conv4} we can pass to the limit as $n\to +\infty$ in \eqref{weak_v} with $v_{n}$ in place of $v$ and we get \begin{equation} \int_{\Omega} \operatorname{\mathbf{a}}(x,v,\nabla u)\nabla \varphi dx =\int_{\Omega}f\varphi dx - \int_{\Omega} \Phi(x,v)\nabla \varphi dx\,, \quad \forall \varphi\in W^{1,p}(\Omega)\,. \notag \end{equation} Since there exists a unique weak solution to \eqref{weak_v} with median equal to zero we obtain that the whole sequence $u_{n}$ converges to $u$ in $L^{p}(\Omega)$ and $u=\Gamma (v)$. It follows that $\Gamma$ is continuous. \par Compactness of $\Gamma $ immediatly follows. Indeed, thank to the assumptions, for any $ v\in L^p(\Omega)$, we have \[ \int_\Omega |\nabla u |^p dx \le C\,, \] where $C$ is a constant depending on $\alpha$, $a_0$, $a_1$, $\|c\|_{L^{\infty}(\Omega)}$, $\Omega$, $N$, $p$ and $f$. Then, using Poincar\'e-Wirtinger inequality and Rellich theorem, $u=\Gamma(v)$ belongs to a compact set of $L^p(\Omega)$. By choosing a ball of $L^p(\Omega)$, $ B_{L^p}(0, r)$ such that \[ \Gamma \left( B_{L^p}(0, r) \right ) \subset B_{L^p}(0, r)\,, \] Leray-Schauder fixed point theorem ensures the existence of at least one fixed point. \end{proof} \section{Existence result for renormalized solutions} In this section we prove our main result which gives the existence of a renormalized solution to problem \eqref{pb}. \begin{theorem} \label{exist_renorm} Assume \eqref{ell}--\eqref{comp}. If the datum $f$ belongs to $ L^{1}(\Omega)$, then there exists at least one renormalized solution $u$ to problem \eqref{pb} having ${\operatorname{med}}(u)=0$. \end{theorem} \begin{proof} The proof is divided into 7 steps. In a standard way we begin by introducing a sequence of approximate problems whose data are smooth enough and converge in some sense to the datum $f$. Then we prove that the weak solutions $u_\varepsilon$ to the approximate problems and their gradients $\nabla u_\varepsilon $ satisfy a priori estimates; such estimates allow to prove that $u_\varepsilon$ and $\nabla u_\varepsilon$ converge to a function $u$ and its gradient $\nabla u$ respectively. The final step consists in proving that $u$ is a renormalized solution to \eqref{pb} by showing that it is possible to pass to the limit in the approximate problems. \par\smallskip \noindent{\sl Step 1. Approximate problems.} \noindent For $\varepsilon >0$, let us define \[\operatorname{\mathbf{a}}_{\varepsilon}(x ,s,\xi)=\operatorname{\mathbf{a}}(x, T_{\frac1\varepsilon}(s), \xi) + \varepsilon|\xi|^{p-2}\xi, \] \[\Phi_{\varepsilon}(x ,s)=T_{\frac1\varepsilon}(\Phi(x, s) ) \] and $f_{\varepsilon}\in L^{p'}(\Omega) $ such that \begin{gather*} \int_{\Omega} f_\varepsilon\, dx = 0, \\ f_{\varepsilon} \rightarrow f\quad \text{strongly in} \>L^1(\Omega), \\ \|f_{\varepsilon}\|_{L^{1}(\Omega)} \leq \|f\|_{L^{1}(\Omega)}\,,\quad \forall \varepsilon>0. \end{gather*} Let us denote by $u_\varepsilon$ one weak solution belonging to $W^{1,p}(\Omega)$ such that \[ \text{med} (u_\varepsilon)=0 \] and \begin{equation} \int_{\Omega} \operatorname{\mathbf{a}}_\varepsilon(x,u_\varepsilon,\nabla u_\varepsilon)\nabla \varphi dx +\int_{\Omega} \Phi_\varepsilon(x,u_\varepsilon)\nabla \varphi dx =\int_{\Omega}f_\varepsilon\varphi dx \, \label{appr_eps} \end{equation} for every $\varphi \in W^{1,p}(\Omega)$. The existence of such a function $u_\varepsilon$ follows from Theorem \ref{exist_weak}. \par \smallskip \noindent {\sl Step 2. A priori estimates} \noindent Using $\varphi =T_k(u_\varepsilon)$ for $k>0$, as test function in \eqref{appr_eps} we have \begin{gather*} \int_{\Omega} \operatorname{\mathbf{a}}_\varepsilon (x,u_\varepsilon,\nabla u_\varepsilon)\nabla T_k(u_\varepsilon) dx +\int_{\Omega} \Phi_\varepsilon(x,T_k (u_\varepsilon))\nabla T_k(u_\varepsilon) dx \\ = \int_{\Omega}f_\varepsilon T_k(u_\varepsilon) dx. \notag \end{gather*} which implies, by \eqref{ell} and \eqref{growthphi}, \begin{align*} \alpha \int_\Omega |\nabla T_k(u_\varepsilon)|^p dx \le \int_\Omega c(x) (1+ |T_k(u_\varepsilon)|^{p-1})|\nabla T_k (u_\varepsilon)| dx +k \| f\|_{L^1(\Omega)}. \notag \end{align*} By Young inequality we get \begin{equation} \label{4eq0} \int_\Omega |\nabla T_k(u_\varepsilon)|^p dx \le M(k+k^p) \end{equation} for a suitable positive constant $M$ which depends on the data, but does not depend on $k$ and $\varepsilon$. \noindent We deduce that, for every $k>0$, \[T_k(u_\varepsilon) \text{ is bounded in } W^{1,p}(\Omega). \] Moreover taking into account \eqref{growth} and \eqref{4eq0}, we obtain that for any $k>0$ \begin{equation*} \operatorname{\mathbf{a}}(x,T_{k}(u_{\varepsilon}),\nabla T_{k}(u_{\varepsilon})) \text{ is bounded in $(L^{p'}(\Omega))^{N}$} \end{equation*} uniformly with respect to $\varepsilon$. Therefore there exists a measurable function $\,u :\Omega \rightarrow \overline\mathbb R$ and for any $k>0$ there exists a function $\sigma_{k}$ belonging to $(L^{p'}(\Omega))^{N}$ such that, up to a subsequence still indexed by $\varepsilon$, \begin{gather} \label{rin_conv1} u_\varepsilon \rightarrow u \text{ a.e. in } \Omega, \\ \label{rin_conv2} T_k(u_\varepsilon) \rightharpoonup T_k(u) \text{ weakly in } W^{1,p}(\Omega), \\ \label{rin_conv3} \operatorname{\mathbf{a}} (x,T_{k}(u_\varepsilon),\nabla T_K(u_\varepsilon))\rightharpoonup \sigma_k \text{ weakly in } (L^{p'}(\Omega))^N \quad\forall k>0. \end{gather} \noindent {\sl Step 3. The function $u$ is finite a.e. in $\Omega$ and ${\operatorname{med}}(u)=0$.} Since ${\operatorname{med}}(u_{\varepsilon})=0$, Poincar\'e-Wirtinger inequality allows us to use a log-type estimate (see \cite{BeGu,BOP,Droniou00,DV} for similar non coercive problems). We consider the function \begin{equation} \Psi_p(r)=\int_0^r\frac{1}{(1+|s|)^p} ds\,, \quad \forall r\in \mathbb R. \notag \end{equation} We observe that ${\operatorname{med}}(\Psi_p(u_\varepsilon))={\operatorname{med}}(u_\varepsilon)=0$. Using $\Psi_p(u_\varepsilon)$ as test function in \eqref{appr_eps}, we get \begin{align*} \int_{\Omega} \operatorname{\mathbf{a}}_\varepsilon(x,u_\varepsilon, \nabla u_\varepsilon) \frac {\nabla u_\varepsilon}{(1+|u_\varepsilon|)^p} dx +\int_{\Omega} \Phi_\varepsilon (x, u_\varepsilon) \frac {\nabla u_\varepsilon}{(1+|u_\varepsilon|)^p} dx = \int_\Omega f_\varepsilon\Psi_p(u_\varepsilon) dx. \end{align*} By ellipticity condition \eqref{ell}, growth condition \eqref{growthphi} and since $\displaystyle \| \Psi_p(u_\varepsilon)\|_{L^{\infty}(\Omega)} \le \frac{1}{p-1}$, we get \begin{align*} \alpha\int_{\Omega} \frac {|\nabla u_\varepsilon|^p}{(1+|u_\varepsilon|)^p}dx &\le \int_{\Omega} c(x) (1+|u_\varepsilon|^{p-1} )\frac {|\nabla u_\varepsilon|}{(1+|u_\varepsilon|)^p} dx + \frac{1}{p-1}\|f\|_{L^{1}(\Omega)}\\ & \le C\int_{\Omega} c(x) \frac {|\nabla u_\varepsilon|}{(1+|u_\varepsilon|)} dx + \frac{1}{p-1}\|f\|_{L^{1}(\Omega)}, \end{align*} where $C$ is a generic and positive constant independent of $\varepsilon$. By Young inequality we deduce \begin{equation} \int_{\Omega} \frac {|\nabla u_\varepsilon|^p}{(1+|u_\varepsilon|)^p}dx \leq Cp' \|c\|_{L^{p'}(\Omega)}^{p'}+ \frac{p'}{p-1}\|f\|_{L^1(\Omega)} \label{stima}. \end{equation} Let us define \[ \Psi_1(u_\varepsilon)=\int_0^{u_\varepsilon}\frac{1}{(1+|s|)}ds={\rm sign}\,(u_{\varepsilon}) \ln(1+|u_{\varepsilon}|) . \] By \eqref{stima} we have \[ \|\nabla \Psi_1(u_\varepsilon)\|_{(L^p(\Omega))^N}\le C \] and since ${\operatorname{med}}(\Psi_1(u_\varepsilon))=0$, Poincar\'e-Wirtinger inequality leads to \[ \|\Psi_1(u_\varepsilon)\|_{L^p(\Omega)}\le C. \] According to the definition of $\Psi_{1}$ we obtain that \begin{equation}\label{eqog0} \sup_{\varepsilon>0} \operatorname{meas}(\{x\in\Omega\,;\,|u_{\varepsilon}(x)|>A\})\leq \frac{C}{\ln(1+A)} \end{equation} and this implies that $u$ is finite almost everywhere in $\Omega$. \par Since ${\operatorname{med}}(u_\varepsilon)=0$ for any $\varepsilon>0$ we also have, for any $k>0$, ${\operatorname{med}}(T_{k}(u_\varepsilon))=0$, for any $\varepsilon>0$. Due to the point-wise convergence of $u_\varepsilon$ and to the fact that $T_{k}(u)\in W^{1,p}(\Omega)$ we obtain that ${\operatorname{med}}(T_{k}(u))=0$ for any $k>0$. It follows that ${\operatorname{med}}(u)=0$. \par\smallskip \noindent{\sl Step 4. }{ \sl We prove} \begin{equation} \label{4eq00} \lim_{n\rightarrow +\infty}\limsup_{\varepsilon\rightarrow 0} \frac{1}{n} \int_{\Omega} \operatorname{\mathbf{a}}_{\varepsilon}(x,u_{\varepsilon},\nabla u_{\varepsilon}) \nabla T_{n}(u_{\varepsilon}) dx = 0. \end{equation} Using the test function $\frac{1}{n}T_{n}(u_{\varepsilon})$ in \eqref{appr_eps} we have \begin{gather*} \frac{1}{n}\int_{\Omega} \operatorname{\mathbf{a}}_\varepsilon (x,u_\varepsilon,\nabla u_\varepsilon)\nabla T_n(u_\varepsilon) dx +\frac{1}{n}\int_{\Omega} \Phi_\varepsilon(x,T_n (u_\varepsilon))\nabla T_n(u_\varepsilon) dx \\ =\frac{1}{n} \int_{\Omega}f_\varepsilon T_n(u_\varepsilon) dx\,, \notag \end{gather*} which yields that \begin{equation} \label{eqog1} \begin{split} \frac{1}{n}\int_{\Omega} \operatorname{\mathbf{a}}_\varepsilon (x,u_\varepsilon,\nabla u_\varepsilon) & \nabla T_n(u_\varepsilon) dx \leq \frac{1}{n} \int_{\Omega}|f_\varepsilon|\times |T_n(u_\varepsilon)| dx \\ & {} + \frac{1}{n}\int_{\Omega} c(x) (1+|T_n (u_\varepsilon)|^{p-1}) |\nabla T_n(u_\varepsilon)| dx. \end{split} \end{equation} Due to \eqref{rin_conv1} the sequence $T_{n}(u_\varepsilon)$ converges to $T_{n}(u)$ as $\varepsilon$ goes to zero in $L^{\infty}(\Omega)$ weak-*. Since $f_{\varepsilon}$ strongly converges to $f$ in $L^{1}(\Omega)$ it follows that \begin{equation*} \lim_{\varepsilon\rightarrow 0}\frac{1}{n} \int_{\Omega}|f_\varepsilon|\times |T_n(u_\varepsilon)| dx= \frac{1}{n} \int_{\Omega}|f|\times |T_n(u)| dx. \end{equation*} Recalling that $u$ is finite almost everywhere in $\Omega$, the sequence $T_{n}(u)/n$ converges to 0 as $n$ goes to infinity in $L^{\infty}(\Omega)$ weak-*. Therefore we deduce that \begin{equation} \label{eqog2} \lim_{n\rightarrow+\infty}\lim_{\varepsilon\rightarrow 0}\frac{1}{n} \int_{\Omega}|f_\varepsilon|\times |T_n(u_\varepsilon)| dx= 0. \end{equation} If $R$ is a positive real number which will be chosen later, let us define for any $\varepsilon>0$ the set $E_{\varepsilon,R}=\{ x\in \Omega\,:\, |u_\varepsilon(x)| > R\}$. We have for any $n>R$ \begin{equation}\label{eqog3} \begin{split} \frac{1}{n}\int_{\Omega} c(x) & (1+|T_n (u_\varepsilon)|^{p-1}) |\nabla T_n(u_\varepsilon)| dx \\ \leq {}& \frac{1}{n}\int_{\Omega\setminus E_{\varepsilon,R}} c(x) (1+| T_{R}(u_\varepsilon)|^{p-1}) |\nabla T_{R}(u_\varepsilon)| dx \\ & {} + \frac{1}{n}\int_{ E_{\varepsilon,R}} c(x) (1+|T_n (u_\varepsilon)|^{p-1}) |\nabla T_n(u_\varepsilon)| dx. \end{split} \end{equation} H\"older inequality yields that \begin{equation*} \begin{split} \frac{1}{n}\int_{\Omega\setminus E_{\varepsilon,R}} & c(x) (1+| T_{R}(u_\varepsilon)|^{p-1}) |\nabla T_{R}(u_\varepsilon)| dx \leq \frac{1+R^{p-1}}{n} \int_{\Omega} c(x) |\nabla T_R(u_\varepsilon)| dx \\ & {} \leq \frac{1+R^{p-1}}{n} \| c\|_{L^{p'}{(\Omega)}} \|\nabla T_R(u_\varepsilon) \|_{(L^p(\Omega))^N} \end{split} \end{equation*} and since $T_R(u_\varepsilon)$ is bounded in $W^{1,p}(\Omega)$ uniformly with respect to $\varepsilon$ we obtain \begin{equation}\label{eqog4} \lim_{n\rightarrow+\infty} \limsup_{\varepsilon\rightarrow 0} \frac{1}{n}\int_{\{|u_{\varepsilon}|\leq R\}} c(x) (1+| T_{R}(u_\varepsilon)|^{p-1}) |\nabla T_{R}(u_\varepsilon)| dx =0. \end{equation} To control the second term of the right-hand side of \eqref{eqog3} we distinguish the case $p<N$ and $p=N$. If $p<N$ we have \[ \frac{p-1}{N}+\frac{(N-p)(p-1)}{Np}+\frac{1}{p}=1 \] so that H\"older inequality gives \begin{equation*} \begin{split} \frac{1}{n}\int_{E_{\varepsilon,R}} & c(x) (1+|T_n (u_\varepsilon)|^{p-1}) |\nabla T_n(u_\varepsilon)| dx \leq \frac{1}{n} \| c\|_{L^{N/(p-1)}(E_{\varepsilon,R})} \\ & \times \Big( \operatorname{meas}(\Omega)^{Np/((N-p)(p-1))} + \| T_n(u_\varepsilon) \|_{L^{pN/(N-p)}(\Omega)} \Big) \| \nabla T_n(u_\varepsilon) \|_{(L^p(\Omega))^N}. \end{split} \end{equation*} Recalling that ${\operatorname{med}}(T_n(u_\varepsilon))=0$ Poincar\'e-Wirtinger inequality and Sobolev embedding theorem lead to \begin{equation}\label{eqog5} \begin{split} \frac{1}{n}\int_{E_{\varepsilon,R}} c(x) & (1+|T_n (u_\varepsilon)|^{p-1}) |\nabla T_n(u_\varepsilon)| dx\\ & \leq \frac{C}{n} \| c\|_{L^{N/(p-1)}(E_{\varepsilon,R})} \Big( 1 + \| \nabla T_n(u_\varepsilon) \|_{(L^p(\Omega))^N}^p \Big) \end{split} \end{equation} where $C>0$ is a constant independent of $n$ and $\varepsilon$. If $p=N$, since $c$ belongs to $L^q(\Omega)$ with $q>\frac{N}{N-1}$ similar arguments lead to \begin{equation}\label{eqog6} \begin{split} \frac{1}{n}\int_{E_{\varepsilon,R}} c(x) &(1+|T_n (u_\varepsilon)|^{p-1}) |\nabla T_n(u_\varepsilon)| dx \\ & \le \frac{C}{n} \| c\|_{L^{q}(E_{\varepsilon,R})} \Big( 1 + \| \nabla T_n(u_\varepsilon) \|_{(L^p(\Omega))^N}^p \Big) \end{split} \end{equation} where $C>0$ is a constant independent of $n$ and $\varepsilon$. \par In view of \eqref{eqog0} and the equi-integrability of $c$ in $L^{q} (\Omega)$ (with $q=N/(p-1)$ if $p<N$ and $q>N/(N-1)$ if $p=N$) let $R>0$ such that for any $\varepsilon>0$ \begin{equation}\label{eqog6b} C \| c\|_{L^{q}(E_{\varepsilon,R})} < \frac{\alpha}{2}\,, \end{equation} where $\alpha$ denotes the ellipticity constant in \eqref{ell}. Using the ellipticity condition \eqref{ell} together with \eqref{eqog1}--\eqref{eqog6b} leads to \[ \frac{1}{n}\int_{\Omega} \operatorname{\mathbf{a}}_\varepsilon (x,u_\varepsilon,\nabla u_\varepsilon) \nabla T_n(u_\varepsilon) dx \leq \frac{C}{n} \|c\|_{L^q(\Omega)}+\omega(\varepsilon,n) \] with $q=N/(p-1)$ if $p<N$ and $q>N/(N-1)$ if $p=N$ and where $\omega(\varepsilon,n)$ is such that $\lim_{n\rightarrow\infty}\limsup_{\varepsilon\rightarrow0} \omega(\varepsilon,n)=0$. \par It follows that \eqref{4eq00} holds. \par\smallskip \noindent{\sl Step 5. We prove that for any $k>0$} \begin{equation} \label{4eq1} \begin{split} \lim_{\varepsilon \rightarrow 0} \int_{\Omega} (\operatorname{\mathbf{a}}(x,T_{k}(u_\varepsilon)&,\nabla T_{k}(u_\varepsilon))-\operatorname{\mathbf{a}}(x,T_{k}(u_\varepsilon),\nabla T_{k}(u))) \\ & \cdot (\nabla T_{k}(u_{\varepsilon})-\nabla T_{k}(u)) dx =0. \end{split} \end{equation} Let $h_{n}$ defined by \begin{equation} \label{h_n} h_n(s)= \begin{cases} 0 &\quad \text{if } |s|>2n, \\ \displaystyle\frac{2n-|s|}{n} &\quad \text{if } n<|s|\le 2n,\\ 1 &\quad \text{ if } |s|\le n\,. \end{cases} \end{equation} Using the admissible test function $h_{n}(u_{\varepsilon}) (T_{k}(u_{\varepsilon})-T_{k}(u))$ to \eqref{appr_eps} we have \begin{equation} \label{4eq2} \begin{split} \int_{\Omega} h_{n}(u_{\varepsilon}) & \operatorname{\mathbf{a}}(x,u_{\varepsilon},\nabla u_{\varepsilon}) (\nabla T_{k}(u_{\varepsilon})- \nabla T_{k}(u)) dx \\ & {} = A_{k,n,\varepsilon} + B_{k,n,\varepsilon} + C_{k,n,\varepsilon} + D_{k,n,\varepsilon} + E_{k,n,\varepsilon} \end{split} \end{equation} with \begin{gather*} A_{k,n,\varepsilon} =\int_{\Omega} h_{n}(u_\varepsilon) f_{\varepsilon} (T_{k}(u_\varepsilon)-T_{k}(u))dx, \\ B_{k,n,\varepsilon}= - \int_{\Omega} h_{n}(u_\varepsilon) \Phi_{\varepsilon}(x,u_\varepsilon) (\nabla T_{k}(u_\varepsilon)-\nabla T_{k}(u)) dx, \\ C_{k,n,\varepsilon}= - \int_{\Omega} h'_{n}(u_\varepsilon) \Phi_{\varepsilon}(x,u_\varepsilon) \nabla u_\varepsilon (T_{k}(u_\varepsilon)-T_{k}(u)) dx, \\ D_{k,n,\varepsilon} = - \int_{\Omega} h'_{n}(u_\varepsilon) \operatorname{\mathbf{a}}_{\varepsilon}(x,u_\varepsilon,\nabla u_\varepsilon) \nabla u_\varepsilon (T_{k}(u_{\varepsilon}) - T_{k}(u))dx, \\ E_{k,n,\varepsilon}= -\varepsilon \int_{\Omega} h_{n}(u_\varepsilon) |\nabla u_\varepsilon |^{p-2} \nabla u_\varepsilon ( \nabla T_{k}(u_{\varepsilon}) -\nabla T_{k}(u))dx. \end{gather*} We now pass to the limit in \eqref{4eq2} first as $\varepsilon$ goes to zero and then as $n$ goes to infinity. \par Due to the point-wise convergence of $u_\varepsilon$ the sequence $T_{k}(u_\varepsilon)-T_{k}(u)$ converges to zero almost everywhere in $\Omega$ and in $L^{\infty}(\Omega)$ weak* as $\varepsilon$ goes to zero. Since $f_{\varepsilon}$ converges to $f$ strongly in $L^{1}(\Omega)$ we obtain that \begin{equation*} \lim_{\varepsilon \rightarrow 0} A_{k,n,\varepsilon} = \lim_{\varepsilon\rightarrow 0} \int_{\Omega} h_{n}(u_{\varepsilon}) f_{\varepsilon} (T_{k}(u_\varepsilon)-T_{k}(u))dx =0. \end{equation*} For $\varepsilon<1/n$ we have $h_{n}(s)\Phi_{\varepsilon}(x,s)=h_{n}(s)\Phi(x,s)$ for any $s\in\mathbb R$ and a.e. in $\Omega$. Using the point-wise convergence of $u_\varepsilon$ $h_{n}(u_\varepsilon) \Phi_{\varepsilon}(x,u_\varepsilon)$ converges to $h_{n}(u)\Phi(x,u)$ a.e. in $\Omega$ as $\varepsilon$ goes to zero while by \eqref{growthphi} we have $h_{n}(u_{\varepsilon})|\Phi_{\varepsilon}(x,u_\varepsilon)|\leq (1+(2n)^{p-1}) c(x)$. It follows that $h_{n}(u_{\varepsilon})\Phi_{\varepsilon}(x,u_\varepsilon)$ converges to $h_{n}(u)\Phi(x,u)$ strongly in $(L^{q}(\Omega))^{N}$ with $q=N/(p-1)$ if $N>p$ and $q>N/(N-1)$ if $N=p$. Due to \eqref{rin_conv2} we deduce that \begin{equation*} \lim_{\varepsilon\rightarrow 0} B_{k,n,\varepsilon}= - \lim_{\varepsilon\rightarrow 0} \int_{\Omega} h_{n}(u_\varepsilon) \Phi_{\varepsilon}(x,u_\varepsilon) (\nabla T_{k}(u_\varepsilon)-\nabla T_{k}(u)) dx =0. \end{equation*} With arguments already used we also have for any $n\geq 1/\varepsilon$ \begin{equation*} \lim_{\varepsilon\rightarrow 0} C_{k,n,\varepsilon}= - \lim_{\varepsilon\rightarrow 0} \int_{\Omega} h'_{n}(u_\varepsilon) \Phi_{\varepsilon}(x,u_\varepsilon) \nabla u_\varepsilon (T_{k}(u_\varepsilon)-T_{k}(u)) dx =0. \end{equation*} Since \begin{equation*} |D_{k,n,\varepsilon}| \leq \frac{2k}{n} \int_{\{|u_\varepsilon|\leq 2n\}} \operatorname{\mathbf{a}}_{\varepsilon}(x,u_\varepsilon,\nabla u_\varepsilon) \nabla u_\varepsilon dx \end{equation*} and due to \eqref{4eq00} we obtain that \begin{equation*} \lim_{n\rightarrow 0}\limsup_{\varepsilon\rightarrow 0 } D_{k,n,\varepsilon}=0. \end{equation*} The identification $h_{n}(u_\varepsilon) |\nabla u_\varepsilon |^{p-2} \nabla u_\varepsilon = h_{n}(u_\varepsilon) |\nabla T_{2n}(u_\varepsilon) |^{p-2} \nabla T_{2n}(u_\varepsilon)$ a.e. in $\Omega$ and estimate \eqref{4eq0} imply that \[ h_{n}(u_\varepsilon) |\nabla u_\varepsilon |^{p-2} \nabla u_\varepsilon ( \nabla T_{k}(u_{\varepsilon}) -\nabla T_{k}(u)) \] is bounded in $L^{1}(\Omega)$ uniformly with respect to $\varepsilon$. It follows that \[ \varepsilon h_{n}(u_\varepsilon) |\nabla u_\varepsilon |^{p-2} \nabla u_\varepsilon ( \nabla T_{k}(u_{\varepsilon}) -\nabla T_{k}(u)) \] converges to 0 strongly in $L^1(\Omega)$ so that \begin{equation*} \lim_{\varepsilon\rightarrow 0 } E_{k,n,\varepsilon}=0. \end{equation*} As a consequence we obtain that for any $k>0$ \begin{equation*} \lim_{n\rightarrow \infty}\limsup_{\varepsilon\rightarrow 0 } \int_{\Omega} h_{n}(u_{\varepsilon}) \operatorname{\mathbf{a}}(x,u_{\varepsilon},\nabla u_{\varepsilon}) (\nabla T_{k}(u_{\varepsilon})- \nabla T_{k}(u)) dx= 0. \end{equation*} Recalling that for any $n>k$, we have \[ h_{n}(u_{\varepsilon}) \operatorname{\mathbf{a}}(x,u_{\varepsilon},\nabla u_{\varepsilon}) \nabla T_{k}(u_{\varepsilon})= \operatorname{\mathbf{a}}(x,u_{\varepsilon},\nabla u_{\varepsilon}) \nabla T_{k}(u_{\varepsilon}) \quad\text{a.e. in $\Omega$.} \] It follows that \begin{multline} \label{4eq3} \limsup_{\varepsilon\rightarrow 0 }\int_{\Omega} \operatorname{\mathbf{a}}(x,u_{\varepsilon},\nabla u_{\varepsilon}) \nabla T_{k}(u_{\varepsilon}) dx \\ \leq \lim_{n\rightarrow \infty}\limsup_{\varepsilon\rightarrow 0 } \int_{\Omega} h_{n}(u_{\varepsilon}) \operatorname{\mathbf{a}}(x,u_{\varepsilon},\nabla u_{\varepsilon}) \nabla T_{k}(u) dx. \end{multline} According to the definition of $h_{n}$ we have \[ h_{n}(u_\varepsilon) \operatorname{\mathbf{a}}(x,u_{\varepsilon},\nabla u_{\varepsilon}) = h_{n}(u_\varepsilon) \operatorname{\mathbf{a}}(x,T_{2n}(u_{\varepsilon}),\nabla T_{2n}(u_{\varepsilon})) \text{ a.e. in $\Omega$} \] so that \eqref{rin_conv1} and \eqref{rin_conv3} give \begin{equation} \label{4eq4} \lim_{\varepsilon\rightarrow 0 } \int_{\Omega} h_{n}(u_{\varepsilon}) \operatorname{\mathbf{a}}(x,u_{\varepsilon},\nabla u_{\varepsilon}) \nabla T_{k}(u) dx = \int_{\Omega} h_{n}(u) \sigma_{2n} \nabla T_{k}(u) dx. \end{equation} \par If $n>k$ we have \[ \operatorname{\mathbf{a}}(x,T_{n}(u_{\varepsilon}),\nabla T_{n}(u_{\varepsilon})) \chi_{\{|u_\varepsilon|<k\}} = \operatorname{\mathbf{a}}(x,T_{k}(u_{\varepsilon}),\nabla T_{k}(u_{\varepsilon})) \chi_{\{|u_\varepsilon|<k\}} \] almost everywhere in $\Omega$. From \eqref{rin_conv1} and \eqref{rin_conv3} it follows that \[ \sigma_{n} \chi_{\{|u|<k\}} = \sigma_{k} \chi_{\{|u|<k\}} \text{ a.e. in $\Omega\setminus\{|u|=k\}$} \] and then we obtain for any $n>k$ \[ \sigma_{n} \nabla T_{k}(u)= \sigma_{k} \nabla T_{k}(u) \text{ a.e. in $\Omega$.} \] Therefore \eqref{4eq3} and \eqref{4eq4} allow to conclude that \begin{equation} \label{4eq5} \limsup_{\varepsilon \rightarrow 0} \int_{\Omega} \operatorname{\mathbf{a}}(x,T_{k}(u_\varepsilon),\nabla T_{k}(u_\varepsilon)) \nabla T_{k}(u_{\varepsilon}) dx \leq \int_{\Omega} \sigma_{k} \nabla T_{k}(u) dx. \end{equation} We are now in a position to prove \eqref{4eq1}. Indeed the monotone character of $\operatorname{\mathbf{a}}$ implies that for any $\varepsilon>0$ \begin{multline} \label{4eq6} 0 \leq \int_{\Omega} (\operatorname{\mathbf{a}}(x,T_{k}(u_\varepsilon),\nabla T_{k}(u_\varepsilon)) -\operatorname{\mathbf{a}}(x,T_{k}(u_\varepsilon),\nabla T_{k}(u))) \\ \cdot (\nabla T_{k}(u_{\varepsilon})-\nabla T_{k}(u)) dx.\qquad\quad \end{multline} Moreover, using the point-wise convergence of $T_{k}(u_\varepsilon)$ and assumption \eqref{growth}, the function $\operatorname{\mathbf{a}}(x,T_{k}(u_\varepsilon),\nabla T_{k}(u))$ converges to $\operatorname{\mathbf{a}}(x,T_{k}(u),\nabla T_{k}(u))$ strongly in $(L^{p'}(\Omega))^{N}$. Writing \begin{equation*} \begin{split} \int_{\Omega} (\operatorname{\mathbf{a}}(x,T_{k}(u_\varepsilon), & \nabla T_{k}(u_\varepsilon))-\operatorname{\mathbf{a}}(x,T_{k}(u_\varepsilon),\nabla T_{k}(u))) (\nabla T_{k}(u_{\varepsilon})-\nabla T_{k}(u)) dx \\ {} = {}& \int_{\Omega} \operatorname{\mathbf{a}}(x,T_{k}(u_\varepsilon),\nabla T_{k}(u_\varepsilon)) (\nabla T_{k}(u_{\varepsilon})-\nabla T_{k}(u)) dx \\ & {} - \int_{\Omega} \operatorname{\mathbf{a}}(x,T_{k}(u_\varepsilon),\nabla T_{k}(u)) (\nabla T_{k}(u_{\varepsilon})-\nabla T_{k}(u)) dx, \end{split} \end{equation*} using \eqref{4eq5} and \eqref{4eq6} allow to conclude that \eqref{4eq1} holds for any $k>0$. \par\smallskip \noindent{\sl Step 6. We prove in this step that for any $k>0$} \begin{gather} \operatorname{\mathbf{a}}(x,T_{k}(u),\nabla T_{k}(u)) = \sigma_{k} \label{4eq7} \\ \operatorname{\mathbf{a}}(x,T_{k}(u_\varepsilon), \nabla T_{k}(u_\varepsilon)) \nabla T_{k}(u_\varepsilon) \rightharpoonup \operatorname{\mathbf{a}}(x,T_{k}(u),\nabla T_{k}(u))\nabla T_{k}(u) \label{4eq8} \end{gather} weakly in $L^{1}(\Omega)$ as $\varepsilon$ goes to zero. \par From \eqref{4eq6} we have for any $k>0$ \begin{equation*} \lim_{\varepsilon \rightarrow 0} \int_{\Omega} \operatorname{\mathbf{a}}(x,T_{k}(u_\varepsilon),\nabla T_{k}(u_\varepsilon)) \nabla T_{k}(u_{\varepsilon}) dx = \int_{\Omega} \sigma_{k} \nabla T_{k}(u) dx. \end{equation*} The monotone character of $\operatorname{\mathbf{a}}$ and the usual Minty argument imply \eqref{4eq7}. \par From \eqref{4eq1} we get \[ (\operatorname{\mathbf{a}}(x,T_{k}(u_\varepsilon), \nabla T_{k}(u_\varepsilon))-\operatorname{\mathbf{a}}(x,T_{k}(u_\varepsilon),\nabla T_{k}(u))) (\nabla T_{k}(u_{\varepsilon})-\nabla T_{k}(u)) \rightarrow 0 \] strongly in $L^{1}(\Omega)$ as $\varepsilon$ goes to zero. Using \eqref{rin_conv2} and recalling that the sequence $\operatorname{\mathbf{a}}(x,T_{k}(u_\varepsilon),\nabla T_{k}(u)))$ converges to $\operatorname{\mathbf{a}}(x,T_{k}(u),\nabla T_{k}(u)))$ strongly in $(L^{p'}(\Omega))^{N}$ the monotone character of $\operatorname{\mathbf{a}}$ leads to \eqref{4eq8}. \par\smallskip \noindent {\sl Step 7. We are now in a position to pass to the limit in the approximated problem. } \par Let $h$ be a function in $W^{1,\infty}(\mathbb R)$ with compact support, contained in the interval $[-k,k]$, $k>0$ and let $\varphi\in W^{1,p}(\Omega)\cap L^{\infty}(\Omega)$. Using $\varphi h(u_\varepsilon)$ as a test function in the approximated problem we have \begin{multline} \int_{\Omega} h(u_\varepsilon) \operatorname{\mathbf{a}}_{\varepsilon} (x,u_\varepsilon, \nabla u_\varepsilon) \nabla \varphi dx + \int_{\Omega} h'(u_\varepsilon) \operatorname{\mathbf{a}}_{\varepsilon} (x,u_\varepsilon, \nabla u_\varepsilon) \nabla u_\varepsilon \varphi dx \label{rin_eps}\\ + \int_{\Omega} h(u_\varepsilon)\Phi_\varepsilon (x, u_\varepsilon) \nabla \varphi dx + \int_{\Omega} h'(u_\varepsilon)\Phi_\varepsilon (x, u_\varepsilon) \nabla u_\varepsilon \varphi dx \\ = {} \int_\Omega f_\varepsilon\varphi h(u_\varepsilon) dx. \end{multline} We want to pass to the limit in this equality. Since $\text{supp}\, h$ is contained in the interval $[-k,k]$, by the strong converge of $f_\varepsilon$ to $f$ and \eqref{rin_conv1} we immediatly obtain \[ \lim_{\varepsilon\to0} \int_\Omega f_\varepsilon\varphi h(u_\varepsilon) dx= \int_\Omega f\varphi h(u) dx. \] Moreover by growth condition \eqref{growthphi} and \eqref{rin_conv1}, using Lebesgue convergence theorem we deduce that \begin{equation} \lim_{\varepsilon\to0}\int_{\Omega} h(u_\varepsilon)\Phi_\varepsilon (x, u_\varepsilon) \nabla \varphi dx=\int_{\Omega} h(u)\Phi (x, u) \nabla \varphi dx.\notag \end{equation} Analogously from \eqref{rin_conv2} we obtain \begin{align*} \lim_{\varepsilon\to0}\int_{\Omega} h'(u_\varepsilon)\Phi_\varepsilon (x, u_\varepsilon) \nabla u_\varepsilon \varphi dx & =\lim_{\varepsilon\to0}\int_{\Omega} h'(u_\varepsilon)\Phi (x, T_{k}(u_\varepsilon)) \nabla T_k(u_\varepsilon ) \varphi dx\\ & = \int_{\Omega} h'(u)\Phi (x, u) \nabla T_k(u) \varphi dx. \end{align*} In view of the definition of $\operatorname{\mathbf{a}}_{\varepsilon}$ and since $\varepsilon |\nabla T_{k}(u_\varepsilon)|^{p-2}\nabla T_{k}(u_\varepsilon)$ converges to zero strongly in $(L^{p'}(\Omega))^{N}$ as $\varepsilon$ goes to zero, {\eqref{rin_conv3}} and \eqref{4eq7} imply that \begin{align*} \lim_{\varepsilon\to0} \int_{\Omega} h(u_\varepsilon) \operatorname{\mathbf{a}}_{\varepsilon}(x,u_\varepsilon, \nabla T_k(u_\varepsilon) ) & \nabla \varphi dx \\ & =\lim_{\varepsilon\to0} \int_{\Omega} h(u_\varepsilon) \operatorname{\mathbf{a}} (x,T_{k}(u_\varepsilon), \nabla T_k(u_\varepsilon) )\nabla \varphi dx \\ & = \int_{\Omega} h(u) \operatorname{\mathbf{a}} (x,u, \nabla T_k(u) )\nabla \varphi dx. \end{align*} From \eqref{4eq8} we get \begin{align*} \lim_{\varepsilon\to0} \int_{\Omega} h'(u_\varepsilon) \operatorname{\mathbf{a}}_{\varepsilon} & (x,u_\varepsilon, \nabla u_\varepsilon) \nabla u_\varepsilon \varphi dx \\ & {} = \lim_{\varepsilon\to0} \int_{\Omega} h'(u_\varepsilon) \operatorname{\mathbf{a}} (x,T_{k}(u_\varepsilon) , \nabla T_k(u_\varepsilon)) \nabla T_k(u_\varepsilon)) \varphi dx \\ &{} = \int_{\Omega} h'(u) \operatorname{\mathbf{a}} (x,u, \nabla T_k(u)) \nabla T_k(u) \varphi dx. \notag \end{align*} Therefore by passing to the limit in \eqref{rin_eps} we obtain condition \eqref{def4} in the definition of renormalized solution. The decay of the truncated energy \eqref{def3} is a consequence of \eqref{4eq00} and \eqref{4eq8}. Since $u$ is finite almost everywhere in $\Omega$ and since $T_{k}(u)\in W^{1,p}(\Omega)$ for any $k>0$ we can conclude that $u$ is a renormalized solution to \eqref{pb} and that ${\operatorname{med}}(u)=0$. \end{proof} \section{Stability result and further remarks} This section is devoted to state a stability result and to prove that if the right-hand side $f$ is regular enough, under additional assumptions on $\operatorname{\mathbf{a}}$, then any renormalized solution is also a weak solution. For $\varepsilon>0$ let $f_\varepsilon$ belonging to $L^{1}(\Omega)$ and $\Phi_\varepsilon : \Omega\times \mathbb R \mapsto \mathbb R^{N}$ a Carath\'eodory function. Assume that there exists $c\in L^{q}(\Omega)$ with $q=N/(p-1)$ if $p<N$ and $q>N/(N-1)$ if $p=N$ such that for any $\varepsilon>0$ \begin{equation}\label{stab00} \left\vert \Phi_\varepsilon(x,s)\right\vert \leq c(x)(\left\vert s\right\vert ^{p-1}+1) \end{equation}% for almost everywhere in $\Omega$ and every $s\in\mathbb R$. For any $\varepsilon>0$ let $u_\varepsilon$ be a renormalized solution (having null median) to the problem \begin{equation} \label{stab1} \begin{cases} -{\operatorname{div}}\left( \operatorname{\mathbf{a}}\left( x, u_\varepsilon, \nabla u_\varepsilon\right)+ \Phi_\varepsilon (x,u_\varepsilon) \right) =f_\varepsilon & \text{in}\ \Omega,\\ \left( \operatorname{\mathbf{a}}\left( x, u_\varepsilon, \nabla u_\varepsilon\right)+ \Phi_\varepsilon (x,u_\varepsilon) \right)\cdot\underline n=0 & \text{on}\ \partial \Omega ,% \end{cases}% \end{equation}% where $\operatorname{\mathbf{a}}$ verifies \eqref{ell}--\eqref{growth}. \par Moreover assume that \begin{equation} \int_\Omega f_\varepsilon\, dx=0,\qquad f_{\varepsilon}\rightarrow f\text{\ strongly in }L^{1}(\Omega) \label{stab2} \end{equation}% and for almost every $x$ in $\Omega$ \begin{equation}\label{stab3} \begin{cases} \Phi_\varepsilon(x,s_{\varepsilon})\rightarrow \Phi(x,s) \\ \text{for every sequence }s_{\varepsilon}\in\mathbb{R}\text{ such that } s_{\varepsilon}\rightarrow \end{cases} \end{equation}% where $\Phi$ is a Carath\'eodory function verifying (as a consequence of \eqref{stab00}) the growth condition \eqref{growthphi}). \par \begin{theorem}\label{thstab} Under the assumptions \eqref{stab00}, \eqref{stab1}, \eqref{stab2}, \eqref{stab3}, up to a subsequence (still indexed by $\varepsilon$) $u_\varepsilon$ converges to $u$ as $\varepsilon$ goes to zero where $u$ is a renormalized solution to \eqref{pb} with null median. More precisely we have \begin{gather} u_{\varepsilon}\rightarrow u\text{ a.e in }\Omega, \label{stab4} \\ \operatorname{\mathbf{a}}(x,T_{k}(u_\varepsilon),\nabla T_{k}(u_\varepsilon)) \nabla T_{k}(u_\varepsilon) \rightharpoonup \operatorname{\mathbf{a}}(x,T_{k}(u),\nabla T_{k}(u)) \nabla T_{k}(u) \label{stab5} \end{gather} weakly $L^{1}(\Omega)$. \end{theorem} \begin{proof}[Sketch of proof] We mainly follow the arguments developed in the proof of Theorem \ref{exist_renorm}. As usual, the crucial point is to obtain {\em a priori} estimates, i.e. \begin{gather}\label{stab6a} T_{k}(u_\varepsilon) \text{ bounded in } W^{1,p}(\Omega), \\ \operatorname{\mathbf{a}}(x,T_{k}(u_\varepsilon),\nabla T_{k}(u_\varepsilon)) \text{ bounded in } (L^{p'}(\Omega))^{N}\text{ for any $k>0$} \label{stab6b} \end{gather} and \begin{gather} \label{stab6d} \lim_{n\rightarrow +\infty}\limsup_{\varepsilon\rightarrow 0} \frac{1}{n} \int_{\Omega} \operatorname{\mathbf{a}}(x,u_{\varepsilon},\nabla u_{\varepsilon}) \nabla T_{n}(u_{\varepsilon}) dx = 0. \end{gather} Even if $T_{k}(u_{\varepsilon })$ is not an admissible test function in the renormalized formulation (see Definition \ref{defrenorm}) it is well known that it can be achieved through the following process. Using $h=h_{n}$, where $h_n$ is defined in \eqref{h_n}, and $\varphi= T_{k}(u_\varepsilon)$ in the renormalized formulation \eqref{def4} we have, for any $n>0$ and any $k>0$ \begin{multline} \label{5eq1} \int_{\Omega} h_{n}(u_\varepsilon) \operatorname{\mathbf{a}} (x,u_\varepsilon, \nabla u_\varepsilon) \nabla T_{k}(u_\varepsilon) dx \\ + \int_{\Omega} h_{n}'(u_\varepsilon) \operatorname{\mathbf{a}} (x,u_\varepsilon, \nabla u_\varepsilon) \nabla u_\varepsilon T_{k}(u_\varepsilon) dx + \int_{\Omega} h_{n}(u_\varepsilon)\Phi_\varepsilon (x, u_\varepsilon) \nabla T_{k}(u_\varepsilon) dx \\ + \int_{\Omega} h_{n}'(u_\varepsilon)\Phi_\varepsilon (x, u_\varepsilon) \nabla u_\varepsilon T_{k}(u_\varepsilon) dx = \int_\Omega f_\varepsilon T_{k}(u_\varepsilon) h_{n}(u_\varepsilon) dx. \end{multline} We now pass to the limit as $n$ goes to infinity. In view of the definition of $h_{n}$ for any $n>k$ we have \begin{equation*} \int_{\Omega} h_{n}(u_\varepsilon) \operatorname{\mathbf{a}} (x,u_\varepsilon, \nabla u_\varepsilon) \nabla T_{k}(u_\varepsilon) dx = \int_{\Omega} \operatorname{\mathbf{a}} (x,u_\varepsilon, \nabla u_\varepsilon) \nabla T_{k}(u_\varepsilon) dx \end{equation*} and \begin{equation*} \int_{\Omega} h_{n}(u_\varepsilon)\Phi_\varepsilon (x, u_\varepsilon) \nabla T_{k}(u_\varepsilon) dx= \int_{\Omega} \Phi (x, u_\varepsilon) \nabla T_{k}(u_\varepsilon) dx . \end{equation*} Since $u_\varepsilon$ is finite almost everywhere in $\Omega$, the function $h_{n}(u_\varepsilon)$ converges to $1$ in $L^{\infty}(\Omega)$ weak*, so that \begin{equation*} \lim_{n\rightarrow +\infty} \int_\Omega f_\varepsilon T_{k}(u_\varepsilon) h_{n}(u_\varepsilon) dx = \int_\Omega f_\varepsilon T_{k}(u_\varepsilon) dx. \end{equation*} Due to \eqref{def3}, we get \begin{equation*} \lim_{n\rightarrow +\infty} \int_\Omega h_{n}'(u_\varepsilon) \operatorname{\mathbf{a}} (x,u_\varepsilon, \nabla u_\varepsilon) \nabla u_\varepsilon T_{k}(u_\varepsilon) dx = 0. \end{equation*} It remains to control the behavior of the forth term to the right hand side of \eqref{5eq1}. Since we have \begin{equation*} \bigg| \int_{\Omega} h_{n}'(u_\varepsilon)\Phi_\varepsilon (x, u_\varepsilon) \nabla u_\varepsilon T_{k}(u_\varepsilon) dx \bigg| \leq \frac{k}{n} \int_{\Omega} |\Phi_\varepsilon(x,u_\varepsilon)|\times |\nabla T_{2n}(u_\varepsilon)| dx \end{equation*} recalling \eqref{ermk1} we obtain that \begin{equation*} \lim_{n\rightarrow+\infty} \bigg| \int_{\Omega} h_{n}'(u_\varepsilon)\Phi_\varepsilon (x, u_\varepsilon) \nabla u_\varepsilon T_{k}(u_\varepsilon) dx \bigg| =0 . \end{equation*} It follows that passing to the limit as $n$ goes to infinity in \eqref{5eq1} leads to \begin{multline} \label{stab6} \int_{\Omega} \operatorname{\mathbf{a}} (x,u_\varepsilon, \nabla u_\varepsilon) \nabla T_{k}(u_\varepsilon) dx + \int_{\Omega}\Phi (x, u_\varepsilon) \nabla T_{k}(u_\varepsilon) dx\\ = \int_\Omega f_\varepsilon T_{k}(u_\varepsilon) dx \end{multline} and then assumptions on $\operatorname{\mathbf{a}}$, $\Phi_\varepsilon$ and $f_\varepsilon$ give \eqref{stab6a} and \eqref{stab6b}. For the same reasons following Step 2 in the proof of Theorem \ref{exist_renorm}, there exists a function $u$ such that, up to a subsequence still indexed by $\varepsilon$, \begin{gather*} \label{stab_conv1} u_\varepsilon \rightarrow u \text{ a.e. in } \Omega, \\ \label{stab_conv2} T_k(u_\varepsilon) \rightharpoonup T_k(u) \text{ weakly in } W^{1,p}(\Omega), \\ \label{stab_conv3} \operatorname{\mathbf{a}} (x,T_{k}(u_\varepsilon),\nabla T_k(u_\varepsilon))\rightharpoonup \sigma_k \text{ weakly in } (L^{p'}(\Omega))^N \quad\forall k>0, \end{gather*} where $\sigma_{k}$ belongs to $L^{p'}(\Omega)$ for any $k>0$. \par Using a similar process to one used to obtain \eqref{stab6} we get \begin{multline}\label{stab7} \int_{\Omega} \operatorname{\mathbf{a}}(x,u_\varepsilon, \nabla u_\varepsilon) \frac {\nabla u_\varepsilon}{(1+|u_\varepsilon|)^p} dx +\int_{\Omega} \Phi_\varepsilon (x, u_\varepsilon) \frac {\nabla u_\varepsilon}{(1+|u_\varepsilon|)^p} dx\\ = \int_\Omega f_\varepsilon\Psi_p(u_\varepsilon) dx, \end{multline} where $\Psi_{p}(r)=\int_{0}^{r}\frac{1}{(1+|s|)^{p}} ds$. Therefore the arguments developed in Steps 3 and 4 imply that $u$ is finite almost everywhere in $\Omega$ and lead to \eqref{stab6d}. Because the sequel of the proof uses mainly admissible test function in the renormalized formulation and the monotone character of the operator we can repeat the same arguments to show that $u$ is a renormalized solution to \eqref{pb} with null median. In particular following Steps 5 and 6 (see \eqref{4eq8} in the proof of Theorem \ref{exist_renorm}) allow to obtain that \eqref{stab5} hold. \end{proof} Now we prove that if $\operatorname{\mathbf{a}}(x,r,\xi)$ is a classical Leray-Lions operator verifying \eqref{3eq3} and if $f\in L^{q}$ with $q\leq (p^{*})'$ then any renormalized solution to \eqref{pb} is also a weak solution to \eqref{pb} belonging to $W^{1,p}(\Omega)$. \begin{proposition}\label{prop5} Assume that \eqref{ell}, \eqref{mon}, \eqref{growthphi}, \eqref{comp} and \eqref{3eq3} hold. Let $u$ be a renormalized solution to \eqref{pb} with ${\operatorname{med}}(u)=0$. If $f\in L^{q}(\Omega)$ with $q\leq (p^{*})'$ if $N>p$ and $q<+\infty$ if $N=p$ then $u$ belongs to $W^{1,p}(\Omega)$ and \begin{gather*} \int_{\Omega} \operatorname{\mathbf{a}}(x,u,\nabla u)\nabla v dx +\int_{\Omega} \Phi(x,u) \nabla v dx =\int_{\Omega} fv dx \end{gather*} for any $v\in W^{1,p}(\Omega)$. \end{proposition} \begin{proof} Let $u$ be a renormalized solution to \eqref{pb}. We can proceed as in the proof of Theorem \ref{thstab} and we obtain \eqref{stab6}. Then we have \begin{equation} \label{5eq2} \int_{\Omega} \operatorname{\mathbf{a}} (x,u, \nabla u) \nabla T_{k}(u) dx + \int_{\Omega}\Phi (x, u) \nabla T_{k}(u) dx = \int_\Omega f T_{k}(u) dx. \end{equation} Using \eqref{ell}, \eqref{growthphi} and the regularity of $f$ we obtain \begin{equation*} \alpha \int_{\Omega} |\nabla T_{k}(u)|^{p} dx \leq \int_{\Omega} c(x) (1+|u|^{p-1}) |\nabla T_{k}(u)| dx + \|f\|_{L^{q}(\Omega)} \|T_{k}(u)\|_{L^{q'}(\Omega)}. \end{equation*} Let $R>0$ be a real number which will be chosen later and denote \begin{equation*} E_{R}=\{ x\in\Omega\,;\, |u(x)|>R\}. \end{equation*} Using again ${\operatorname{med}}(T_{k}(u))=0$, Poincar\'e-Wirtinger inequality and Sobolev embedding Theorem we have \begin{equation*} \begin{split} \alpha \int_{\Omega} |\nabla T_{k}(u)|^{p} dx \leq & \int_{\Omega} c(x) |\nabla T_{k}(u)| dx + \int_{E_{R}} c(x) |u|^{p-1} |\nabla T_{k}(u)| dx \\ & {} + \int_{\Omega\setminus E_{R}} c(x) |u|^{p-1} |\nabla T_{k}(u)| dx + \|f\|_{L^{q}(\Omega)} \|T_{k}(u)\|_{L^{q'}(\Omega)} . \end{split} \end{equation*} If follows that \begin{equation*} \begin{split} \int_{\Omega} |\nabla T_{k}(u)|^{p} dx \leq {} & C \Big( \| \nabla T_{k}(u)\|_{(L^{p}(\Omega))^{N}} + \|c\|_{L^{q}(E_{R})} \| \nabla T_{k}(u)\|_{(L^{p}(\Omega))^{N}}^{p} \\ & {} + R^{p-1} \| \nabla T_{R}(u)\|_{(L^{p}(\Omega))^{N}}\Big) \end{split} \end{equation*} where $C>0$ depends on $\alpha$, $f$, $N$, $p$, $\operatorname{meas}(\Omega)$, $c$ but is independent of $k$. Since $u$ is finite a.e. in $\Omega$, $\lim_{R\rightarrow +\infty}\operatorname{meas}(E_{R})=0$. By the equi-integrability of $c$ in $L^{q}(\Omega)$ we can choose $R>0$ such that $C \|c\|_{L^{q}(E_{R})}$ is sufficiently small enough so that \begin{equation*} \int_{\Omega} |\nabla T_{k}(u)|^{p} dx \leq C \Big( \| \nabla T_{k}(u)\|_{(L^{p}(\Omega))^{N}} + R^{p-1} \| \nabla T_{R}(u)\|_{(L^{p}(\Omega))^{N}}\Big) \end{equation*} where $C>0$ does not depend on $k$. It follows that \begin{equation*} \int_{\Omega} |\nabla T_{k}(u)|^{p} dx \leq C \end{equation*} where $C>0$ depends on $\alpha$, $f$, $N$, $p$, $\Omega$, $c$, $R$ but is independent of $k$. Since ${\operatorname{med}}(T_k(u))=0$ Poincar\'e-Wirtinger inequality implies that $T_k(u)$ is bounded in $W^{1,p}(\Omega)$ uniformly with respect to $k$. Therefore we conclude that $u$ belongs to $W^{1,p}(\Omega)$. \par Using the renormalized formulation \eqref{def4} with $h=h_{n}$ and passing to the limit as $n$ goes to infinity leads to \begin{gather}\label{5eq3} \int_{\Omega} \operatorname{\mathbf{a}}(x,u,\nabla u)\nabla v dx +\int_{\Omega} \Phi(x,u) \nabla v dx =\int_{\Omega} fv dx \end{gather} for any $v\in L^{\infty}(\Omega)\cap W^{1,p}(\Omega)$. Due to growth assumptions \eqref{3eq3} on $\operatorname{\mathbf{a}}$ and \eqref{growthphi} on $\Phi$ we deduce that $\operatorname{\mathbf{a}}(x,u,\nabla u)$ and $\Phi(x,u)$ belong to $(L^{p'}(\Omega))^{N}$. It follows that \eqref{5eq3} holds for any $v\in W^{1,p}(\Omega)$. \end{proof} \section{Operator with a zero order term} In this section we consider Neumann problems which are similar to \eqref{pb0} with a zero order term. Precisely let us consider the following Neumann problem \begin{equation} \label{newpb} \begin{cases} \lambda(x,u) -{\operatorname{div}}\left( \operatorname{\mathbf{a}}\left( x, u, \nabla u\right)+ \Phi (x,u) \right) =f & \text{in}\ \Omega, \\ \left( \operatorname{\mathbf{a}}\left( x, u, \nabla u\right)+ \Phi (x,u) \right)\cdot\underline n=0 & \text{on}\ \partial \Omega% \end{cases}% \end{equation}% where $\lambda : \Omega\times \mathbb R$ is a Carath\'eodory function verifying \begin{gather} \label{zeroorder0} \lambda(x,s)s \geq 0, \\ \label{zeororder1} \forall k>0,\ \exists c_{k}>0\text{ such that } |\lambda(x,s)|\leq c_{k} \quad\forall |s|\leq k,\text{ a.e. in }\Omega, \\ \label{zeroorder2} \forall s\in\mathbb R\quad |\lambda(x,s)|\geq g(s) \text{ a.e. in $\Omega$} \end{gather} where $g$ is function such that $\lim_{s\rightarrow \pm \infty}g(s)= +\infty$. \par If $f$ belongs to $L^{1}(\Omega)$ and without additional growth assumptions on $g$ we cannot expect to have in general a solution (in whatever sense) lying in $L^{1}(\Omega)$ and then we have similar difficulties to deal with \eqref{newpb}. In particular the presence of $\lambda(x,u)$ does not help to deal with the term $-{\operatorname{div}}(\Phi(x,u))$ and we cannot follow the approach of \cite{AMST97,Droniou00,DV,Prignet97} which use the mean value. However the ``median'' tool and some modifications of the proof of Theorem \ref{exist_renorm} allow to show that there exists at least a renormalized solution to \eqref{newpb}: \begin{theorem} \label{exist_renorm_zero_order_term} Assume \eqref{ell}--\eqref{comp} and \eqref{zeroorder0}--\eqref{zeroorder2}. If the datum $f$ belongs to $ L^{1}(\Omega) $ then there exists at least one renormalized solution $u$ to problem \eqref{newpb}. \end{theorem} \begin{proof}[Sketch of proof] As in Theorem \ref{exist_weak}, a fixed point theorem and classical results of Leray-Lions give the existence of $u_\varepsilon$ belonging to $W^{1,p}(\Omega)$ verifying \begin{multline} \varepsilon \int_{\Omega}|u_\varepsilon|^{p-2} u_\varepsilon v dx + \int_{\Omega}\lambda(x,T_{1/\varepsilon}(u_\varepsilon)) v dx \\ + \int_{\Omega} \operatorname{\mathbf{a}}(x,T_{1/\varepsilon}(u_\varepsilon),\nabla u_\varepsilon)\nabla v dx \\ +\int_{\Omega} \Phi(x,T_{1/\varepsilon}(u_\varepsilon))\nabla v dx =\int_{\Omega}T_{1/\varepsilon}(f) v dx \label{appr_newpb_eps} \end{multline} for any $v$ lying in $W^{1,p}(\Omega)$. Due to the zero order term $\varepsilon |u_\varepsilon|^{p-2} u_\varepsilon +\lambda(x,T_{1/\varepsilon}(u_\varepsilon))$ in the equation, we do not need any compatibility condition on $f$. The counter part is that we cannot expect to have (or to fix) ${\operatorname{med}}(u_\varepsilon)=0$ and then it yields another difficulties. In particular Steps 3 and 4 (see the proof of Theorem \ref{exist_renorm}) which use strongly the fact that the solution has a null median should be adapted in the case of the approximated problem \eqref{appr_newpb_eps}. Step 2 is unchanged and we have the following and additional estimate \begin{equation} \label{neu00} T_{1/\varepsilon}(g(u_\varepsilon)) \text{ bounded in } L^{1}(\Omega). \end{equation} Due to the behavior at infinity of the function $g$ we deduce that \begin{gather} \label{neu01} \lim_{A\rightarrow +\infty}\sup_{\varepsilon>0 } \operatorname{meas}\{ x\in \Omega;\, |u_\varepsilon(x)|>A\} =0, \\ \label{neu02} \forall \varepsilon>0 \quad |{\operatorname{med}}(u_\varepsilon)|\leq M \end{gather} where $M$ is a positive real number independent of $\varepsilon$. It follows (after extracting appropriate subsequence, see Step 2) that there exists a measurable function $u$ which is finite almost everywhere in $\Omega$ such that \begin{gather*} u_\varepsilon \rightarrow u \text{ a.e. in } \Omega, \\ T_k(u_\varepsilon) \rightharpoonup T_k(u) \text{ weakly in } W^{1,p}(\Omega), \quad\forall k>0. \end{gather*} Step 4 which is crucial in dealing with renormalized solutions consists here in proving that \begin{equation} \label{neu03} \lim_{n\rightarrow +\infty}\limsup_{\varepsilon\rightarrow 0} \frac{1}{n} \int_{\Omega} \operatorname{\mathbf{a}}(x,u_{\varepsilon},\nabla u_{\varepsilon}) \nabla T_{n}(u_{\varepsilon}) dx = 0 \end{equation} using the test function $T_{n}(u_\varepsilon)$ in \eqref{appr_newpb_eps}. Due to the sign condition \eqref{zeroorder0} the contribution of the zero order terms \begin{equation*} \varepsilon \int_{\Omega}|u_\varepsilon|^{p-2} u_\varepsilon T_{n} (u_\varepsilon) dx + \int_{\Omega}\lambda(x,T_{1/\varepsilon}(u_\varepsilon)) T_{n}(u_\varepsilon) dx \end{equation*} is positive. It follows that the inequality \eqref{eqog1} holds: \begin{equation*} \begin{split} \frac{1}{n}\int_{\Omega} \operatorname{\mathbf{a}} (x,u_\varepsilon,\nabla u_\varepsilon) & \nabla T_n(u_\varepsilon) dx \leq \frac{1}{n} \int_{\Omega}|T_{1/\varepsilon}(f)|\times |T_n(u_\varepsilon)| dx \\ & {} + \frac{1}{n}\int_{\Omega} c(x) (1+|T_n (u_\varepsilon)|^{p-1}) |\nabla T_n(u_\varepsilon)| dx. \end{split} \end{equation*} Because we do not have in the present case the property ${\operatorname{med}}(u_\varepsilon)={\operatorname{med}}(T_{n}(u_\varepsilon))=0$ we have to modify the estimate of the term \begin{equation}\label{neu04} \frac{1}{n} \int_{\Omega} c(x) (1+|T_{n}(u_\varepsilon)|^{p-1}) |\nabla T_{n}(u_\varepsilon)| dx. \end{equation} In view of \eqref{neu02} we have for any $n>0$ and for any $\varepsilon>0$ $|{\operatorname{med}}(T_{n}(u_\varepsilon))| \leq M$. It follows that by writing $T_{n}(u_\varepsilon)=T_{n}(u_\varepsilon)-{\operatorname{med}}(T_{n}(u_\varepsilon)) + {\operatorname{med}}(T_{n}(u_\varepsilon))$ we obtain \begin{equation}\label{neu05} \begin{split} \frac{1}{n} \int_{\Omega} c(x) &(1+|T_{n}(u_\varepsilon)|^{p-1}) |\nabla T_{n}(u_\varepsilon)| dx \\ & \leq \frac{C}{n} \int_{\Omega} c(x) (1+|T_{n}(u_\varepsilon)-{\operatorname{med}}(T_{n}(u_\varepsilon))|^{p-1}) |\nabla T_{n}(u_\varepsilon)| dx \end{split} \end{equation} where $C>0$ is a constant independent of $\varepsilon$ and $n$. Poincar\'e-Wirtinger inequality \eqref{poincare}, similar arguments to the ones developed in Step 4 and \eqref{neu02} then allow conclude that \eqref{neu03} holds. \par As far as Step 5 is concerned, it is sufficient to remark that the Lebesgue Theorem yields that \begin{gather*} \lim_{n\rightarrow +\infty} \lim_{\varepsilon\rightarrow0 }\varepsilon \int_{\Omega} h_{n}(u_\varepsilon) |u_\varepsilon|^{p-2} u_\varepsilon (T_{k} (u_\varepsilon)-T_{k}(u)) dx =0 \\ \lim_{n\rightarrow +\infty} \lim_{\varepsilon\rightarrow0 } \int_{\Omega} h_{n}(u_\varepsilon) \lambda(x,T_{1/\varepsilon}(u_\varepsilon)) (T_{k} (u_\varepsilon)-T_{k}(u)) dx =0. \end{gather*} Since Step 6 remains unchanged, in Step 7 we pass to the limit as $\varepsilon$ goes to zero in \begin{multline*} \varepsilon \int_{\Omega} h(u_\varepsilon) |u_\varepsilon|^{p-2} u_\varepsilon \varphi dx + \int_{\Omega} h(u_\varepsilon) \lambda(x,T_{1/\varepsilon}(u_\varepsilon)) \varphi dx \\ + \int_{\Omega} h(u_\varepsilon) \operatorname{\mathbf{a}} (x,u_\varepsilon, \nabla u_\varepsilon) \nabla \varphi dx + \int_{\Omega} h'(u_\varepsilon) \operatorname{\mathbf{a}} (x,u_\varepsilon, \nabla u_\varepsilon) \nabla u_\varepsilon \varphi dx \\ + \int_{\Omega} h(u_\varepsilon)\Phi (x, u_\varepsilon) \nabla \varphi dx + \int_{\Omega} h'(u_\varepsilon)\Phi (x, u_\varepsilon) \nabla u_\varepsilon \varphi dx \\ = {} \int_\Omega T_{1/\varepsilon}(f)\varphi h(u_\varepsilon) dx \end{multline*} where $h$ is a Lispchitz continuous function with compact support and where $\varphi$ lies in $W^{1,p}(\Omega)\cap L^{\infty}(\Omega)$. Since the Lebesgue Theorem gives that \begin{gather*} \lim_{\varepsilon\rightarrow 0} \varepsilon \int_{\Omega} h(u_\varepsilon) |u_\varepsilon|^{p-2} u_\varepsilon \varphi dx =0 \\ \lim_{\varepsilon\rightarrow 0} \int_{\Omega} h(u_\varepsilon) \lambda(x,T_{1/\varepsilon}(u_\varepsilon)) \varphi dx = \int_{\Omega} h(u) \lambda(x,u) \varphi dx \end{gather*} the attentive reader may convince by himself that we obtain the existence of a renormalized solution to equation \eqref{newpb}. \end{proof} \section*{Acknowledgement} This work was done during the visits made by the first and the third authors to Laboratoire de Math\'ematiques ``Rapha\"el Salem'' de l'Universit\'e de Rouen and by the second author to Dipartimento di Matematica e Applicazioni ``R. Caccioppoli'' dell' Universit\`a degli Studi di Napoli ``Federico II''. Hospitality and support of all these institutions are gratefully acknowledged.
\section{Introduction} Deep neural networks have been successful at solving many kinds of tasks \cite{Bengio2012}. Parallel processors such as GPUs have played a significant role in the practical implementation of deep neural networks. The computations that arise when training and using deep neural networks lend themselves naturally to efficient parallel implementations. The efficiency provided by these implementations allows researchers to explore significantly higher capacity networks, training them on larger datasets \cite{Coates2013}. This has led to greatly improved accuracy on tasks such as speech recognition and image classification, among others. For example, the majority of entries in the 2014 ILSVRC challenge \cite{ILSVRC2014} use GPUs to implement deep neural networks, as pioneered by \cite{Krizhevsky2012}, and improved by many others, including \cite{Sermanet2013} \cite{Szegedy2014}. Deep neural networks for speech recognition have also benefited from parallel implementations on GPUs \cite{Hinton2012} \cite{Dahl2012} \cite{Maas2014}. Convolutional Neural Networks (CNNs) \cite{Lecun1998} are an important and successful class of deep networks. Convolutional neural networks are computed using dense kernels that differ from traditional dense linear algebra routines. Accordingly, modern deep learning frameworks, such as Torch7 \cite{Collobert2011}, Theano \cite{Bergstra2010}, and Caffe \cite{Jia2014} feature suites of custom kernels that implement basic operations such as tensor convolutions, activation functions and pooling. These routines represent the bulk of the computation when training a CNN, and thus account for the majority of its execution time. The deep learning community has been successful in finding optimized implementations of these kernels, but as the underlying architectures evolve, these kernels must be re-optimized, which is a significant investment. Optimizing these kernels requires a deep understanding of the underlying processor architecture, with careful scheduling of data movement, on-chip memory placement, register blocking, and other optimizations in order to get acceptable performance. We believe that providing a library of optimized routines for these computations will provide several important benefits. Firstly, deep learning frameworks can focus on higher-level issues rather than close optimization of parallel kernels to specific hardware platforms. Secondly, as parallel architectures evolve, library providers can provide performance portability, in much the same way as the BLAS routines provide performance portability to diverse applications on diverse hardware. Thirdly, a clearer separation of concerns allows specialization: library providers can take advantage of their deep understanding of parallel architectures to provide optimal efficiency. Our goal is to make it much easier for deep learning frameworks to take advantage of parallel hardware. Our library addresses this goal by providing a flexible, easy-to-use C-language API for deep learning workloads that integrates neatly into existing frameworks. It can provide immediate efficiency gains, and it is rigorously tested and maintained in order to be reliable and performant across a range of different processor architectures. Importantly, our library is designed to use the minimum possible amount of auxiliary memory, which frees up scarce memory for larger models and datasets. We also optimize performance across a wide range of potential use cases, including small mini-batch sizes. \section{Library} One of the primary goals of cuDNN is to enable the community of neural network frameworks to benefit equally from its APIs. Accordingly, users of cuDNN are not required to adopt any particular software framework, or even data layout. Rather than providing a layer abstraction, we provide lower-level computational primitives, in order to simplify integration with existing deep learning frameworks, each with their own abstractions. The bulk of the API is dedicated to functions that perform primitive operations on data stored in user-controlled buffers. By keeping the API low-level, the library integrates simply into other frameworks. cuDNN supports forward and backward propagation variants of all its routines in single and double precision floating-point arithmetic. These include convolution, pooling and activation functions. The library allows variable data layout and strides, as well as indexing of sub-sections of input images. It also includes a set of auxiliary tensor transformation routines that allow for the easy manipulation of 4d-tensors. \subsection{Overview and Handles} The library exposes a host-callable C language API, but requires that input and output data be resident on the GPU, analogously to cuBLAS. The library is thread-safe and its routines can be called from different host threads. Convolutional routines for the forward and backward passes use a common descriptor that encapsulates the attributes of the layer. Tensors and Filters are represented in opaque descriptors, with the flexibility to specify the tensor layout using arbitrary strides along each dimension for tensors. \subsubsection{Spatial Convolutions} The most important computational primitive in convolutional neural networks is a special form of batched convolution. The parameters governing this convolution are listed in table \ref{tab:conv_param}. In this section, we describe the forward form of this convolution - the other forms necessary for backpropagation are closely related. \begin{table} \centering \begin{tabular}{ll} Parameter & Meaning \\ \hline $N$ & Number of images in mini-batch \\ $C$ & Number of input feature maps \\ $H$ & Height of input image \\ $W$ & Width of input image \\ $K$ & Number of output feature maps \\ $R$ & Height of filter kernel \\ $S$ & Width of filter kernel \\ \hline $u$ & Vertical stride \\ $v$ & Horizontal stride \\ $pad\_h$ & Height of zero-padding \\ $pad\_w$ & Width of zero-padding \\ \end{tabular} \caption{Convolutional parameters} \label{tab:conv_param} \end{table} There are two inputs to the convolution: $D \in \mathbb{R}^{NCHW}$, which forms the input data, and $F \in \mathbb{R}^{KCRS}$, which forms the convolutional filters. The input data ranges over $N$ images in a mini-batch, $C$ input feature maps, $H$ rows per image, and $W$ columns per image. The filters range over $K$ output feature maps, $C$ input feature maps, $R$ rows per filter, and $S$ columns per filter. The output is also a four-dimensional tensor $O \in \mathbb{R}^{NKPQ}$, where $N$ and $K$ were as defined previously, and $P=f(H, R, u, pad\_h)$, $Q=f(W, S, v, pad\_w)$, meaning that the height and width of the output images depends on the image and filter height and width, along with padding and striding choices. The striding parameters $u$ and $v$ allow the user to reduce the computational load by computing only a subset of the output pixels. The padding parameters allow users to specify how many rows or columns of $0$ entries are appended to each image. MATLAB's ``valid'' convolution mode corresponds to setting $pad\_h=0, pad\_w=0$, while MATLAB's ``same'' convolution mode corresponds to $pad\_h=\left\lfloor\frac{R}{2}\right\rfloor, pad\_w=\left\lfloor\frac{S}{2}\right\rfloor$, and MATLAB's ``full'' convolution mode corresponds to $pad\_h=R-1, pad\_w=S-1$. More specifically, \begin{equation} f(H, R, u, pad\_h) = \left\lceil\frac{H - R + 1 + 2pad\_h}{u}\right\rceil \end{equation} Define an accessing function to account for striding, padding, and inverting for the convolution: \begin{equation} g(p, u, R, r, pad\_h) = p\cdot u + R - r - 1 - pad\_h \end{equation} Then, the forward convolution evaluates $O[n, k, p, q] \; \forall n \in [0, N), \forall k \in [0, K), \forall p \in [0, P), \forall q \in[0, Q)$. For convenience, define $D_0$ as a zero-extended version of $D$. \begin{equation} O[n, k, p, q] = \sum_{c=0}^{C-1} \sum_{r=0}^{R-1} \sum_{s=0}^{S-1} F[k, c, r, s] \cdot D_0[n, c, g(p, u, R, r, pad\_h), g(q, v, S, s, pad\_w)] \label{eq:conv} \end{equation} As can be seen from Equation \ref{eq:conv}, computing the convolution involves a seven-way nested loop, with four independent loops and three accumulation loops. There are many ways of implementing this computation, some of which we will discuss in the next section. cuDNN's convolutional routines incorporate implementations of both the convolution as well as the cross-correlation variants of these functions. These functions support user-defined strides along each dimension of the input and output tensors. This is important because different frameworks store tensors using different memory layouts; for example, some frameworks interleave feature maps, while others keep them separate. cuDNN allows the user to specify memory layout, which makes it much simpler to integrate into existing frameworks. cuDNN's routines also have a mode to either return the raw gradients or to accumulate them in a buffer as needed for models with shared parameters or a directed acyclic graph structure. \subsection{Other functions} cuDNN also provides other commonly used functions for deep learning. For example, it provides three commonly used neuron activation functions; Sigmoid, Rectified Linear and Hyperbolic Tangent. It provides a softmax routine, which by default uses the numerically stable approach of scaling each element to avoid overflow in intermediate results. Softmax may be computed per image across the feature map, height and width dimensions or per spatial location, per image across the feature map dimension. cuDNN provides average and max pooling operations, as well as a set of tensor transformation routines, such as those that add tensors, with optional broadcasting. The goal of providing these functions is to reduce the amount of parallel code that is required for deep learning frameworks, by providing flexible, well-optimized versions of these commonly used functions. With cuDNN, it is possible to write programs that train standard convolutional neural networks without writing any parallel code, but simply using cuDNN and cuBLAS. \section{Implementation} The majority of functions that cuDNN provides have straightforward implementations. The convolution implementation is not as obvious, so we will outline the motivation and reasoning behind our design choices. There are several ways to implement convolutions efficiently. Our goal is to provide performance as close as possible to matrix multiplication, while using no auxiliary memory. GPU memory is high bandwidth, but low capacity, and is therefore a scarce resource. When training deep networks, ideally the GPU memory should be filled with data, parameters, and neuron responses, not auxiliary data structures needed by the convolution algorithm. Several approaches to computing convolutions require large auxiliary data structures, and therefore we do not consider these approaches for cuDNN. One approach is to lower the convolutions into a matrix multiplication, following \cite{Chellapilla2006}. This can be done by reshaping the filter tensor $F$ into a matrix $F_m$ with dimensions $K \times CRS$, and gathering a data matrix by duplicating the original input data into a matrix $D_m$ with dimensions $CRS \times NPQ$. The computation can then be performed with a single matrix multiply to form an output matrix $O_m$ with dimension $K \times NPQ$. \begin{figure}[h] \centering \includegraphics[width=4in]{multiconvolve} \caption{Convolution lowering} \label{fig:lower} \end{figure} Figure \ref{fig:lower} illustrates how a simple convolution can be lowered to a matrix multiplication. The colors in this illustration represent the input feature maps, and elements of $D$ and $F$ are uniquely labeled in the illustration so as to show how each participates in forming $D_m$ and $F_m$. The filter matrix $F_m$ has dimensions $K \times CRS = 2 \times 12$, while the data matrix $D_m$ has dimensions $CRS \times NPQ = 12 \times 4$. Note that each element of $D$ is duplicated up to $RS=4$ times in $D_m$. The output matrix $O_m$ has dimensions $K \times NPQ = 2 \times 4$. Lowering convolutions to matrix multiplication can be efficient, since matrix multiplication is highly optimized. Matrix multiplication is fast because it has a high ratio of floating-point operations per byte of data transferred. This ratio increases as the matrices get larger, meaning that matrix multiplication is less efficient on small matrices. Accordingly, this approach to convolution is most effective when it creates large matrices for multiplication. As we mentioned earlier, the sizes of $D_m$ and $F_m$ depend on products of the parameters to the convolution, not the parameters themselves. This means that performance using this approach can be very consistent, since the algorithm does not care if one of the parameters is small, as long as the {\em product} is large enough. For example, it is often true that in early layers of a convolutional network, $C$ is small, but $R$ and $S$ are large, while at the end of the network, $C$ is large, but $R$ and $S$ are small. However, the product $CRS$ is usually fairly large for all layers, so performance can be consistently good. The disadvantage of this approach is that forming $D_m$ involves duplicating the input data up to $RS$ times, which can require a prohibitively large temporary allocation. To work around this, implementations sometimes materialize $D_m$ piece by piece, for example, by calling matrix multiplication iteratively for each element of the mini-batch. However, this limits the parallelism in the implementation, and can lead to cases where the matrix multiplications are too small to effectively utilize the GPU. This approach also lowers the computational intensity of the convolutions, because $D_m$ must be written and read, in addition to reading $D$ itself, requiring significantly more memory traffic as a more direct approach. Accordingly, we opt not to use this implementation directly, although as we will explain, our implementation is related. Another approach is to use the Fast Fourier Transform to compute the convolution. The FFT can significantly lower the work complexity of the convolutions, and with clever engineering can be used effectively for deep neural networks \cite{Mathieu2013}. However, the FFT based approach uses a significant amount of temporary memory, since the filters must be padded to be the same size as the inputs. This is especially costly when the filters are small compared to the images, which often happens in the first few layers of a convolutional network. Additionally, the FFT based approach does not perform efficiently when the striding parameters $u$ and $v$ are greater than $1$, which is common in many state-of-the art networks, such as the early layers in \cite{Sermanet2013} and \cite{Szegedy2014}. Striding reduces the computational work of the convolutions by a factor of $uv$, by computing only a sparse subset of the output. However, the nature of the FFT algorithm is such that computing pruned FFTs is a non-trivial task, and often is slower than computing the dense FFT, followed by an additional subsampling step. Due to these drawbacks, we opted to forgo the FFT approach, although we agree that in some cases it is useful. Another common approach is to compute the convolutions directly. This can be very efficient, but requires a large number of specialized implementations to handle the many corner cases implicit in the 11-dimensional parameter space of the convolutions. Implementations following this approach are often well-optimized for convolutions in certain parts of the parameter space, but perform poorly for others. For example, cuda-convnet2 \cite{cudaconvnet2-2014} performs well when batch sizes are large, but poorly once batch size falls to 64 or below. Optimizing and maintaining all these specializations is a difficult task. As we envision this library being maintained for some time, and being ported to yet-to-be-conceived future architectures, we searched for something simpler that would perform more robustly across the parameter space and be easier to port to new architectures. \subsection{Our approach} NVIDIA provides a matrix multiplication routine that achieves a substantial fraction of floating-point throughput on GPUs. The algorithm for this routine is similar to the algorithm described in \cite{Tan2011}. Fixed sized submatrices of the input matrices $A$ and $B$ are successively read into on-chip memory and are then used to compute a submatrix of the output matrix $C$. We compute on tiles of $A$ and $B$ while fetching the next tiles of $A$ and $B$ from off-chip memory into on-chip caches and other memories. This technique hides the memory latency associated with the data transfer, allowing the matrix multiplication computation to be limited only by the time it takes to perform the arithmetic. As we discussed earlier, convolutions can be lowered onto matrix multiplication. This approach provides simplicity of implementation as well as consistency of performance across the parameter space, although materializing the lowered matrix in memory can be costly. Our solution follows this approach, but we avoid the problems with materializing the lowered matrix in memory by lazily materializing $D_m$ in on-chip memory only, rather than by materializing it in off-chip memory before calling a matrix multiplication routine. Since the tiling required for the matrix multiplication routine is independent of any parameters from the convolution, the mapping between the tile boundaries of $D_m$ and the convolution problem is non-trivial. Accordingly, our approach entails computing this mapping and using it to load the correct elements of $A$ and $B$ into on-chip memories. This happens dynamically as the computation proceeds, which allows our convolution algorithm to exploit optimized infrastructure for matrix multiplication. We require additional indexing arithmetic compared to a matrix multiplication, but fully leverage the computational engine of matrix multiplication to perform the work. After the computation is complete, we perform the required tensor transposition to store the result in the user's desired data layout. Computing the additional indexing requires repeated calculations of integer division and modulus operations by launch-time constant divisors. We make use of the algorithm for integer division and modulus presented in \cite{Warren2003} to transform these costly operations into integer multiplies and shifts, and thus reduce the indexing overhead required by our approach. \subsection{Performance} The convolution routines in cuDNN provide competitive performance with zero auxiliary memory required. \begin{figure} \centering \includegraphics[width=4in]{comparative_performance} \caption{Comparative Performance} \label{fig:comp_perf} \end{figure} \begin{table} \centering \begin{tabular}{llllllllllll} & $N$ & $C$ & $H$ & $W$ & $K$ & $R$ & $S$ & $u$ & $v$ & $pad\_h$ & $pad\_w$ \\ \hline Layer 1 & 128 & 3 & 128 & 128 & 96 & 11 & 11 & 1 & 1 & 0 & 0 \\ Layer 2 & 128 & 96 & 64 & 64 & 128 & 9 & 9 & 1 & 1 & 0 & 0 \\ Layer 3 & 128 & 128 & 32 & 32 & 128 & 9 & 9 & 1 & 1 & 0 & 0 \\ Layer 4 & 128 & 128 & 16 & 16 & 128 & 7 & 7 & 1 & 1 & 0 & 0 \\ Layer 5 & 128 & 128 & 13 & 13 & 384 & 3 & 3 & 1 & 1 & 0 & 0 \end{tabular} \caption{Convolutional layer collection} \label{tab:layers} \end{table} Figure \ref{fig:comp_perf} shows the performance on an NVIDIA Tesla K40 of three convolution implementations: cuDNN, Caffe, and cuda-convnet2. We evaluated these implementations using the layer configurations shown in table \ref{tab:layers}, which are commonly used for benchmarking convolution performance \cite{Soumith2014}, and quote average throughput for all these layers. cuDNN performance ranges from $0.8\times$ to $2.25\times$ that of cuda-convnet2, with an advantage at smaller batch sizes. Compared to Caffe, cuDNN performance ranges from $1.0\times$ to $1.41\times$. Importantly, even with a small mini-batch size of only 16, cuDNN performance is still 86\% of maximum performance, which shows that our implementation performs well across the convolution parameter space. \begin{table} \centering \begin{tabular}{l|ll|ll} & K40 (TFlops) & GTX980 (TFlops) & K40 (\%) & GTX980 (\%) \\ \hline Layer 1 & 0.99 & 1.46 & 23\% & 30\% \\ Layer 2 & 1.52 & 2.49 & 35\% & 51\% \\ Layer 3 & 1.51 & 2.42 & 35\% & 49\% \\ Layer 4 & 1.38 & 2.19 & 32\% & 45\% \\ Layer 5 & 1.01 & 1.87 & 24\% & 38\% \\ \end{tabular} \caption{Performance portability} \label{tab:perf_port} \end{table} Table \ref{tab:perf_port} illustrates cuDNN's performance portability across GPU architectures. The Tesla K40 is built using the Kepler architecture, and has a peak single-precision throughput of 4.29 TFlops. The Geforce GTX 980 is built using the newer Maxwell architecture, and has a peak single-precision throughput of 4.95 TFlops. cuDNN performance ranges from 23-35\% of peak on the Tesla K40, and from 30-51\% of peak on the GTX 980. This data illustrates how cuDNN provides performance portability across GPU architectures, with no need for users to retune their code as GPU architectures evolve. \section{Caffe Integration} Caffe is a deep learning framework developed with expression, speed, and modularity in mind. Its architecture mirrors the natural modularity of deep networks as compositional models made from a collection of inter-connected layers. A deep network is defined layer-by-layer in a plaintext schema. Each layer type is implemented according to a simple protocol of setup, forward, and backward steps to encapsulate engineering details. Data and derivatives flow through layers and between the host and device according to the framework's unified memory interface that handles allocation and communication. cuDNN integration raises the speed and memory efficiency of the framework without sacrificing expression or modularity. \subsection{Development} Integration is made simple by the self-contained design of the cuDNN handles, descriptors, and function calls together with the modularity of the framework. The core Caffe framework is unaltered, preserving the network, layer, and memory interfaces. Changes were isolated to new layer definitions and implementations, helper functions for descriptors, and the corresponding tests. The patch is almost purely additive. In Caffe each type of model operation is encapsulated in a layer. Layer development comprises declaring and implementing a layer class, defining the layer in the protocol buffer model schema, extending the layer factory, and including tests. Computations are carried out through the layer protocol of setup, forward, and backward steps. cuDNN layers fit this same scheme. Library handles and descriptors are configured in setup while forward and backward calls are made in the respective layer methods. The cuDNN primitives yield concise layer implementations. The cuDNN layers are drop-in replacements to their standard Caffe counterparts. Caffe has a standard array and memory interface, called a blob, for storing and communicating data on the host and device. Blobs hold data, gradients, and parameters. In particular, layer inputs and outputs are held in $N \times C \times H \times W$ dimensional blobs. cuDNN tensor and filter descriptors are trivially constructed from blob arrays through its flexible support for dimension and stride. By coordinating memory solely at the descriptor, Caffe retains control over memory and communication for efficiency. The Caffe$+$cuDNN convolution layer exploits the reduced memory consumption of cuDNN to speed up execution. The forward pass parallelizes the computation of group convolution (filtering with sets of filters with restricted channel connectivity). The backward pass parallelizes the computation of the gradients with respect to the bias, filter weights, and bottom data. \begin{table} \centering \begin{tabular}{llll} & Caffe (seconds) & Caffe$+$cuDNN (seconds) & Speedup \\ \hline Backward Propagation & 156270 & 120651 & 1.30 \\ Forward Propagation & 109310 & 75330 & 1.45 \\ Testing & 21600 & 14529 & 1.49 \\ Update & 2926 & 2920 & 1.00 \\ Overall & 290106 & 213431 & 1.36 \end{tabular} \caption{Caffe Performance} \label{tab:caffe} \end{table} Table \ref{tab:caffe} illustrates the performance improvements gained from integrating cuDNN into Caffe. Overall, training time for 200 iterations improved by 36\%, when training the bvlc\_reference\_caffenet model, using cuDNN R1 on an NVIDIA Tesla K40. \subsection{User Experience} cuDNN computation is transparent to the user through drop-in integration. The model schema and framework interfaces are completely unchanged. Setting a single compilation flag during installation equips Caffe with cuDNN layer implementations and sets cuDNN as the default computation engine. Layers automatically fall back to the standard Caffe functionality in all cases outside of the current scope of cuDNN. Model execution can be tuned through per-layer \textsc{engine} parameters that select the implementation. The workflow for the user is virtually identical. Future cuDNN releases will be integrated in like fashion. \section{Baidu Integration} Several deep learning projects at Baidu have integrated cuDNN. For example, it has been integrated into PADDLE, Baidu's internal deep learning framework. On a Tesla K10 GPU, we observed that performance on a benchmark set of convolutional layers improves by 30\% on average over an implementation that lowers convolution to matrix multiply. We are also using cuDNN in other domains besides image processing, such as speech and language. cuDNN's ability to convolve non-square inputs with asymmetric padding is particularly useful for these other domains. In our experience, cuDNN has reduced memory consumption compared with matrix multiplication, which has enabled us to use larger models and larger mini batches. cuDNN was simple to integrate into our code because we didn't need to change our data structure layout, thanks to cuDNN's flexible interface. \section{Future Work} We are considering several avenues for expanding the performance and functionality of cuDNN. Firstly, although our convolution routines are competitive with other available implementations, more work remains to bring performance up to that attained by matrix multiplication. Over time, we hope to shrink this gap. Secondly, we envision adding support for other primitives; for example: 1D and 3D convolutions would be useful for speech, language processing, and video applications, among others. Local Receptive Field computations, which are similar to convolutions, but with untied weights, are also useful and could be added to cuDNN. Finally, we would like this library to help people use multiple GPUs to accelerate training. \section{Conclusion} This paper presents cuDNN, a library for deep learning primitives. We presented a novel implementation of convolutions that provides reliable performance across a wide range of input sizes, and takes advantage of highly-optimized matrix multiplication routines to provide high performance, without requiring any auxiliary memory. We also provide a set of routines that allow users to train and evaluate complete deep neural networks without the need to write parallel code manually. As parallel architectures continue to evolve, such libraries will provide increasing value to the machine learning community. The library is available at \cite{cuDNN2014}, and we welcome feedback at \url{<EMAIL>}. \bibliographystyle{plain}
\section{Introduction: a formal analogy} Analogies between physical systems, either of mathematical or physical nature, often play a fundamental catalyst role in conceptual and/or technical developments of the respective theories~\cite{JonaLasinio:2010rt}. We discuss here a mathematical analogy between the descriptions of black hole horizons and quantum charged particles, that opens a domain of cross-fertilization between quantum mechanics and gravitation theory. More specifically, apparent horizons --namely {\em marginally outer trapped surfaces} (MOTS)-- possess a stability notion that guarantees their physical consistency as models of black hole horizons. Such MOTS-stability notion \cite{AndMarSim05} admits a spectral characterization in terms of the so-called {\em principal} eigenvalue of the operator \begin{eqnarray} \label{e:MOTS_stability_operator} L_{\cal S} = -\Delta + 2 \Omega^a D_a - \left( |\Omega|^2 - D_a \Omega^a -\frac{1}{2}R_{\cal S} + G_{ab}k^a\ell^b \right) \ \ , \end{eqnarray} defined on the apparent horizon ${\cal S}$. The terms modifying the Laplacian $\Delta$ on ${\cal S}$ are determined by the intrinsic and extrinsic geometry of the apparent horizon and the gravitational equations via the Einstein tensor $G_{ab}$ (see next section for details). The relevant remark in the present context is that under the complexification of the vector $\Omega^a$ and the identifications \begin{eqnarray} \label{e:MOTS_QuantumParticle_Analogy} \Omega_a \leftrightarrow \frac{ie}{\hbar c}A_a \ \ , \ \ R_{\cal S} \leftrightarrow \frac{4me}{\hbar^2}\phi \ \ , \ \ G_{ab}k^a\ell^b \leftrightarrow -\frac{2m}{\hbar^2}V \ \ , \end{eqnarray} the MOTS-stability operator becomes $\frac{\hbar^2}{2m}L_{\cal S} \leftrightarrow \hat{H}$, where \begin{eqnarray} \label{e:quantum_Ham} \hat{H} &=& -\frac{\hbar^2}{2m}\Delta + \frac{i\hbar e}{mc} A^a D_a + \frac{i\hbar e}{2mc} D_a A^a + \frac{e^2}{2mc^2} A_a A^a \nonumber \\ &&+ e \phi + V = \frac{1}{2m}\left(-i\hbar D - \frac{e}{c} A\right)^2 + e\phi + V \ \ , \end{eqnarray} is the Hamiltonian of a non-relativistic particle with mass $m$ and charge $e$ moving on ${\cal S}$ under magnetic and electric fields with vector and scalar potentials given by $A^a$ and $\phi$, and an external potential $V$. This formal mathematical analogy relies on a simple but crucial remark: the derivative and $\Omega^a$ terms in $L_{\cal S}$ can be collected in a perfect square as follows \begin{eqnarray} \label{e:MOTS_stability_operator_v2} L_{\cal S} \psi = \left[-\left(D - \Omega\right)^2 +\frac{1}{2}R_{\cal S} - G_{ab}k^a\ell^b \right]\psi \ \ . \end{eqnarray} Beyond its aesthetic appeal, and in spite of the key difference in the self-adjoint nature of the operators, this analogy has the potential to open bridges between the well-studied quantum particle problem and the rich but largely uncharted MOTS subject, with applications ranging from the MOTS-spectral problem to the spinorial formulation of MOTS stability. We focus here on the study of the full $L_{\cal S}$ spectrum, a challenging problem formulated in \cite{Jaramillo:2013rda} in the setting of a black hole/fluid analogy, but with a definite geometric interest on its own. \section{Geometry and stability of MOTS} \subsection{Some MOTS geometry} Let us consider a codimension-$2$ surface ${\cal S}$, spacelike, closed (compact and without boundary) and orientable, embedded in a $n$-dimensional spacetime $(M, g_{ab})$. The spacetime Levi-Civita connection is denoted by $\nabla_a$ with Einstein curvature tensor $G_{ab}$. Let us denote the induced metric on ${\cal S}$ by $q_{ab}$, with Levi-Civita connection $D_a$, Ricci scalar $R_{\cal S}$, volume form $\epsilon_{ab}$ and measure $\eta_{\cal S}$. Let us consider future-oriented null vectors $\ell^a$ and $k^a$ spanning the normal bundle $T^\perp{\cal S}$ and normalized as $\ell^a k_a = -1$. This normalization leaves a null-vector-rescaling freedom by a positive function $f>0$ respecting time orientation, corresponding to a boost transformation \begin{eqnarray} \label{e:gauge_freedom} \ell'^a =f \ell^a \ \ , \ \ k'^a = f^{-1} k^a \ . \end{eqnarray} We define the expansion $\theta^{(\ell)}$ and H\'aji\v{c}ek{} or rotation form $\Omega_a$ \begin{eqnarray} \label{e:theta_Omega} \theta^{(\ell)}\equiv q^{ab}\nabla_a\ell_b \ \ \ , \ \ \ \Omega_a\equiv -k^c {q^d}_a \nabla_d \ell_c \ , \end{eqnarray} associated with $\ell^a$. Considering $v^a = \alpha \ell^a + \beta k^b$, we can write ${q^c}_a\nabla_c v_b = D_a^\perp v_b + \Theta_{ab}^{(v)}$ , with $\Theta_{ab}^{(v)}\equiv{q^c}_a {q^d}_b \nabla_c v_d$ and \begin{eqnarray} \label{e:normalconnection1} D_a^\perp v_b = (D_a \alpha + \Omega_a \alpha )\ell_b + (D_a \beta - \Omega_a \beta )k_b \ . \end{eqnarray} The H\'aji\v{c}ek{} form therefore provides a connection on the normal bundle $T^\perp {\cal S}$ for the tangent derivative of normal vectors. From a physical perspective it represents a sort of angular momentum density. Given an axial Killing vector $\phi^a$ on ${\cal S}$, \begin{eqnarray} \label{e:angular_momentum} J[\phi]=\frac{1}{8\pi}\int_{\cal S} \Omega_a \phi^a \eta_{\cal S} \end{eqnarray} is the (Komar) angular momentum associated with ${\cal S}$. Regarding the expansion, the surface ${\cal S}$ is a marginally outer trapped surface (MOTS) if it satisfies the condition: $\theta^{(\ell)}=0$. We refer then to $\ell^a$ as {\em outgoing} and to $k^a$ as {\em ingoing} null vectors. \subsection{MOTS stability and MOTS-stability operator} A MOTS ${\cal S}$ is said to be {\em stable} (more properly, stably outermost \cite{AndMarSim05,AndMarSim07}) in the ingoing $k^a$ direction if it can be infinitesimally deformed along $k^a$ into a properly (outer) trapped surface ${\cal S}'$, i.e. with $\theta^{(\ell)}|_{{\cal S}'} < 0$. Using the deformation operator $\delta_v$ along a normal vector $v^a$ (discussed in \cite{AndMarSim05}), this amounts to the existence of a function $\psi>0$ such that $\delta_{\psi k}\theta^{(\ell)}<0$. Such a condition admits a spectral characterization in terms of the elliptic operator $L_{\cal S}$ defined as $L_{\cal S} \psi \equiv \delta_{\psi (-k)} \theta^{(\ell)}$, with explicit expression (\ref{e:MOTS_stability_operator}). The operator $L_{\cal S}$, namely the MOTS-stability operator, is generically non-selfadjoint [in $L^2({\cal S},\eta_{\cal S})]$ due to the $2\Omega^a D_a$ term. Therefore, in the eigenvalue problem \begin{eqnarray} \label{e:MOTS_spectral_problem} L_{\cal S} \psi = \lambda \psi \ , \end{eqnarray} the $\lambda$'s are generically complex. Their real part is bounded below, leading to the definition of the {\em principal eigenvalue} $\lambda_o$ of $L_{\cal S}$ as that with smallest real part. Lemmas 1 and 2 in \cite{AndMarSim05} state that: i) $\lambda_o$ is real, and ii) ${\cal S}$ is stably outermost iff $\lambda_o \geq 0$. \subsection{MOTS-gauge symmetry} The MOTS geometry described above does not depend on the choice of null normals (subject to $\ell^ak_a=-1$). In this sense, the null vector rescaling freedom (\ref{e:gauge_freedom}) is a gauge transformation of the MOTS geometry. We consider now the transformation of the main objects on ${\cal S}$ under the rescaling (\ref{e:gauge_freedom}). {\bf Lemma 1} (MOTS-gauge transformations). {\em Under the null normal rescaling $\ell'^a= f \ell^a$, $k'^a= f^{-1} k^a$, with $f>0$: \begin{itemize} \item[i)] The expansion and H\'aji\v{c}ek{} form transform as \begin{eqnarray} \label{e:theta_Omega_transformations} \theta^{(\ell')}=f \theta^{(\ell)} \ \ \ , \ \ \ \Omega'_a = \Omega_a + D_a(\mathrm{ln}f) \ . \end{eqnarray} \item[ii)] The MOTS-stability operator transforms covariantly \begin{eqnarray} \label{e:transformation_L_S} (L_{\cal S})' \psi = f L_{\cal S} (f^{-1}\psi) \ , \end{eqnarray} where $(L_{\cal S})'\psi \equiv \delta_{\psi (-k')} \theta^{(\ell')}$. \item[iii)] The MOTS-eigenvalue problem is invariant under the additional eigenfunction transformation, $\psi'= f\psi$ \begin{eqnarray} \label{e:transformation_spectral_problem} L_{\cal S}\psi = \lambda \psi \ \ \to \ \ (L_{\cal S})' \psi' = \lambda \psi' \ . \end{eqnarray} \end{itemize} } {\em Proof:} Point {\em i)} follows directly by plugging (\ref{e:gauge_freedom}) into (\ref{e:theta_Omega}). Regarding point {\em ii)}, although it can be obtained by straightforward substitution of (\ref{e:theta_Omega_transformations}) into (\ref{e:MOTS_stability_operator}), it is simpler to use the definition of $L_{\cal S}$. Considering its action on a function $\psi$ \begin{eqnarray} \label{e:L_k_transformation_2} (L_{\cal S})'\psi &=& \delta_{\psi (-k')} \theta^{(\ell')} = \delta_{\psi (-k')} (f\theta^{(\ell)}) \\ &=&\delta_{\psi (-k')} (f)\theta^{(\ell)}+ f \delta_{\psi (-k')} \theta^{(\ell)} = f \delta_{\psi (-k')} \theta^{(\ell)} \nonumber \\ &=& f \delta_{\psi (-f^{-1}k)} (\theta^{(\ell)}) = f \delta_{(f^{-1}\psi) (-k)} \theta^{(\ell)} = f L_{\cal S} (f^{-1}\psi) \nonumber \end{eqnarray} where in the first line we have used the $\theta^{(\ell)}$ transformation in (\ref{e:theta_Omega_transformations}), the second line uses the Leibnitz rule holding for $\delta_v$ and the MOTS condition, and in the third line we used again the definition of $L_{\cal S}$. Finally, point {\em iii)} follows directly \begin{eqnarray} (L_{\cal S})'\psi' = f L_{\cal S} (f^{-1}\psi') = f L_{\cal S} (\psi) = f \lambda \psi = \lambda \psi' \ . \end{eqnarray} \medskip Point {\em i)} just states the invariance of the MOTS notion under (\ref{e:gauge_freedom}) and the transformation of $\Omega_a$ as a connection under the (multiplicative) abelian gauge group $\mathbb{R}^+$ of positive null rescalings. The latter is consistent with the nature of $\Omega_a$ in (\ref{e:normalconnection1}) as a connection in the normal bundle. Point {\em ii)}, stating the good (covariant) transformation properties of $L_{\cal S}$ under $\mathbb{R}^+$, is the analogue in the present setting of Proposition 4 in \cite{Mars:2012sb} concerning the free choice of section of stationary black hole horizons. Point {\em iii)} guarantees that the MOTS-eigenvalue problem is well-defined and provides the gauge transformation rule for the associated eigenfunctions. Of course, all these points evoke familiar features of the quantum charged particle. \section{MOTS and quantum charged particles} To take a step further from the formal correspondance (\ref{e:MOTS_QuantumParticle_Analogy}) into a more precise statement, let us review the stationary quantum charged particle (QCP) problem. The Schr\"odinger equation for a non-relativistic (spin-$0$) charged particle moving in electromagnetic fields with magnetic vector potential $A^a$ and electric potential $\phi$, namely $i\hbar \partial_t \Psi = \hat{H}\Psi$ [with $\hat{H}$ in (\ref{e:quantum_Ham})], follows from that of a non-charged particle in an external mechanical potential $V$ via a minimal-coupling prescription \begin{eqnarray} \label{e:minimal_coupling} i\hbar \partial_t \to i\hbar \partial_t - e \phi \ \ , \ \ -i\hbar D_a \to -i\hbar D_a -\frac{e}{c}A_a \ . \end{eqnarray} The stationary equation for the energy eigenvalues $E$ is then \begin{eqnarray} \label{e:stationary_Schrodinger} \left[\frac{1}{2m}(-i\hbar D_a -\frac{e}{c}A_a)^2 + e\phi + V\right]\psi = E \psi \ , \end{eqnarray} where $\Psi = e^{-iEt/\hbar} \psi$ (with $\partial_t \psi = 0$). This equation should not depend on the gauge choice of the electromagnetic potentials. The gauge transformation of $A_a$ by a total gradient~\footnote{The electric potential $\phi$ stays invariant $\phi \to \phi - \frac{1}{c}\partial_t \sigma$ under gauge transformations compatible with stationarity, $\partial_t \sigma=0$.} \begin{eqnarray} \label{e:electromagnetic_gauge_transformations} A_a \to A_a - D_a \sigma \ , \end{eqnarray} leaves Eq. (\ref{e:stationary_Schrodinger}) invariant if we simultaneously transform $\psi$ as \begin{eqnarray} \label{e:wave_function_transformation} \psi \to e^{ie\sigma/(c\hbar)} \psi \ \ , \end{eqnarray} i.e. by a (local) phase. Transformations (\ref{e:electromagnetic_gauge_transformations}) and (\ref{e:wave_function_transformation}) define the electromagnetic abelian $U(1)$-gauge symmetry. From these remarks we can state the following similarities between the eigenvalue problems (\ref{e:MOTS_spectral_problem}) and (\ref{e:stationary_Schrodinger}), placing the MOTS-QCP analogy in (\ref{e:MOTS_QuantumParticle_Analogy}) on a sounder structural basis~\footnote{ A further point could be the understanding of MOTS-stability, $\lambda_o\geq 0$, as a MOTS-counterpart of a positivity condition on the quantum ground state $E_o$, refining quantum stability. This is however delicate, since the operator correspondance (\ref{e:MOTS_QuantumParticle_Analogy}) does not necessarily preserve eigenvalue signs (see section \ref{s:MOTSspectralproblem}). }: {\em i) Abelian gauge symmetry}. The QCP eigenvalue problem (\ref{e:stationary_Schrodinger}) and the MOTS-spectral problem (\ref{e:MOTS_spectral_problem}) are respectively invariant under transformations (Eqs. (\ref{e:electromagnetic_gauge_transformations})-(\ref{e:wave_function_transformation}) and Lemma 1) \begin{equation} \label{e:compared_gauge_transformations} \begin{array}{rcclcl} \hbox{QCP:} & A_a &\to& A_a - D_a \sigma & , & \psi \to e^{ie\sigma/(c\hbar)} \psi \\ \hbox{MOTS:} & \Omega_a &\to& \Omega_a - D_a \sigma & , & \psi \to e^{-\sigma} \psi \ .\\ \end{array} \end{equation} They both define abelian symmetries of gauge nature: in the QCP case it is the electromagnetic $U(1)$-gauge transformation ($g''= g\cdot g'$, with $g=e^{ie\sigma/(c\hbar)}$) relying on the phase invariance of the wave function, whereas for MOTSs it defines a non-compact $\mathbb{R}^+$-gauge counterpart ($g''= g\cdot g'$, with $g\!=\!f=\!e^{-\sigma}$) reflecting the in-built null rescaling (boost) freedom of the MOTS geometric description. In brief, the MOTS-spectral problem presents symmetry transformation properties in full analogy with those of the QCP Schr\"odinger equation. {\em ii) Gauge field potential}. In this symmetry setting, the 1-form $\Omega_a$ emerges as the natural gauge field of the $\mathbb{R}^+$-gauge group. This endorses, at a structural level, its purely formal correspondence in (\ref{e:MOTS_QuantumParticle_Analogy}) with the $A_a$ magnetic $U(1)$-gauge field. We note that the normal connection in (\ref{e:normalconnection1}) admits an interpretation as a gauge connection: $\ell^b D_a^\perp (\psi k_b) = -(D_a -\Omega_a)\psi$. {\em iii) Minimal Coupling.} It is at a ``dynamical'' level where the analogy proves remarkable: the $\Omega_a$ field enters in $L_{\cal S}$ via a standard gauge ``minimal coupling'' mechanism, namely a shift in the Levi-Civita connection with the gauge connection \begin{eqnarray} \label{e:minimal_coupling_Omega} D_a \to D_a - \Omega_a \ . \end{eqnarray} This becomes apparent in the perfect-square version (\ref{e:MOTS_stability_operator_v2}) of $L_{\cal S}$. Therefore, in full analogy with the minimal coupling mechanism for incorporating the magnetic field in the QCP problem via the shift (\ref{e:minimal_coupling}) in the non-charged equations, rotation in a MOTS is switched-on via the minimal coupling (\ref{e:minimal_coupling_Omega}). \subsection{MOTSs and a negative ``fine structure constant'' $\alpha$} Setting $\hbar\!\!=\!\!m\!\!=\!\!c\!\!=\!\!1$ and introducing a formal complex ``fine-structure constant'' $\alpha\equiv e^2$, we define the operator family \begin{eqnarray} \label{e:L[alpha]} \!\!\!\!\!\! &&L[\sqrt{\alpha}] = -\frac{1}{2}(D -i\sqrt{\alpha}\Omega)^2 - \frac{\alpha}{4}R_{\cal S} - \frac{1}{2}G_{ab}k^a\ell^b \\ \!\!\!\!\!\!&=& \frac{-1}{2}\Delta + i\sqrt{\alpha}(\Omega\cdot D + \frac{1}{2}D\cdot \Omega) + \frac{\alpha}{2}|\Omega|^2 - \frac{\alpha}{4}R_{\cal S} - \frac{1}{2}G(k,\ell). \nonumber \end{eqnarray} The QCP Hamiltonian corresponds to the (normalized) standard real positive $\alpha=1$, whereas (half) the MOTS-stability operator corresponds to a negative $\alpha=-1$. Specifically, QCP and MOTS operators are recovered with branch choices: $\hat{H}=L[\sqrt{\alpha}=1]$ and $L_{\cal S}/2=L[\sqrt{\alpha}=-i]$. In this sense, stable MOTSs can be seen as QCPs with negative ``fine-structure constant'' $\alpha$. This suggests to import QCP terms to MOTSs. {\em Terminology for $L_{\cal S}$ terms.} Regarding terms containing the rotation field $\Omega_a$, we refer to $|\Omega|^2$ as the {\em diamagnetic} term, whereas $\Omega\cdot D$ is the {\em paramagnetic} term \cite{GalPasII91}. The divergence $D\cdot\Omega$ is a {\em gauge-fixing} term and can be chosen by an appropriate transformation (\ref{e:compared_gauge_transformations}). For completeness sake, $\Delta$ is the {\em kinematical} term and $G(k,\ell)$ is the {\em external mechanical} potential. Finally, $R_{\cal S}/4$ can be referred to as the {\em electric} potential term. To justify its explicit distinction from the {\em external mechanical} potential, we consider in the $2$-dimensional case the complex scalar ${\mathcal K}$ on ${\cal S}$ introduced by Penrose and Rindler \cite{PenRin86} as \begin{eqnarray} {\mathcal K} = \frac{1}{4}R_{\cal S} + i \frac{1}{4}\epsilon^{ab}F^\Omega_{ab} \ , \end{eqnarray} where $F^\Omega_{ab}=D_a\Omega_b -D_b\Omega_a$, namely the curvature of $\Omega_a$. The real and imaginary parts of ${\mathcal K}$ correspond, respectively, to {\em electric} and {\em magnetic} terms. This gravity/electromagnetic analogy in ${\mathcal K}$ has been used to discuss isolated/dynamical horizon source multipoles \cite{AshEngPaw04} and to introduce the notions of ``vortexes'' and ``tendexes'' in the analysis of dynamical black holes \cite{OweBriChe11}. The present discussion promotes such analogy to a sounder structural level by identifying the symmetry and minimal coupling similarities in the relevant operators. The ultimate motivation behind this analogy is to explore the transfer of concepts and tools between both problems. This can prove fruitful for the MOTS-spectral problem by profiting of the extensive knowledge accumulated about QCP bound states. In parallel, the development of spinor treatments of MOTS-stability can largely benefit from the presented analogy. In particular, using the Lichnerowicz-Weitzenb\"ock formula to mimic Pauli's approach to spin can provide insights in the structure of the MOTS second-order operator $L_{\cal S}$, whereas Dirac's approach to spin can open an avenue to a first-order formulation of MOTS-stability (of potential interest in boundary value problems of bulk spinor equations). The GHP formalism \cite{gerochheld73:_ghp} can offer additional insights in this setting. We postpone the development of spinor approaches to future studies and focus in the following on the MOTS-spectral problem. \section{MOTS-spectral problem} \label{s:MOTSspectralproblem} \subsection{An explicit example: ``MOTS-Landau'' levels} \label{s:explicitexample} We consider now the behaviour of the eigenvalues and eigenfunctions in problems (\ref{e:stationary_Schrodinger}) and (\ref{e:MOTS_spectral_problem}), under the complex rotation (from $\sqrt{\alpha}=1$ to $\sqrt{\alpha}=-i$, in $L[\sqrt{\alpha}]$) that realizes the analogy (\ref{e:MOTS_QuantumParticle_Analogy}). In this context, and following the spirit of Landau levels of a QCP moving in a constant magnetic field (that provides an explicit example illustrating basic features of such quantum systems), we start by discussing a simple eigenvalue problem for $L_{\cal S}$ that can be explicitly solved both in the MOTS case and, independently, in the analogous QCP case. Let us take a 2-sphere ${\cal S}=S^2$ with ``round'' metric $q_{ab}=r^2(d\theta^2 + \sin^2 \theta d\varphi^2)$. Decomposing the H\'aji\v{c}ek{} form as $\Omega_a = {\epsilon_a}^b D_b \omega + D_a \zeta$ with the simplest non-trivial choice $\omega= a \sin \theta, \zeta =0$ ($a\in \mathbb{R}$), we have: $\Omega = a \sin^2 \theta d\varphi$. In vacuum (i.e. $G_{ab}=8\pi T_{ab}=0$), the solution to the MOTS-eigenvalue problem for the resulting $L_{\cal S}$ is given explicitly by \begin{eqnarray} \label{e:Landau_MOTS} \!\!\!\!\!\!\!\!\!\!\!\!\! \lambda = \frac{1}{r^2}\left[(\lambda_{\ell m}(a) + 1 - a^2) + i 2 a m\right] , \psi = S_{\ell m}(a,\cos\theta) e^{im\varphi} \end{eqnarray} where $S_{\ell m}(a,\cos\theta)$ are the {\em prolate} spheroidal functions with eigenvalues $\lambda_{\ell m}(a)$ (21.6.2 in \cite{AbramowitzStegun64}). Standard spherical harmonics are recovered in the limit $a\to 0$: $\lambda_{\ell m}(a)\to \ell(\ell+1)$ and $S_{lm}(a,\cos\theta) \to P_{\ell m}(\cos \theta)$). We can now consider the ``QCP counterpart'' by performing the complex rotation $a\to i a$ in the operator $L_{\cal S}$. The resolution of the new eigenvalue problem leads to eigenvalues $\bar{\lambda}$ and eigenfunctions $\bar{\psi}$ \begin{eqnarray} \label{e:Landau_QCP} \!\!\!\!\!\!\!\!\!\!\!\!\!\bar{\lambda} = \frac{1}{r^2}\left(\bar{\lambda}_{\ell m}(a) + 1 + a^2 - 2 a m \right) \ , \ \bar{\psi} = \bar{S}_{\ell m}(a,\cos\theta) e^{im\varphi} \end{eqnarray} where $\bar{S}_{\ell m}(a,\cos\theta)=S_{\ell m}(ia,\cos\theta)$ are now the {\em oblate} spheroidal functions with eigenvalues $\bar{\lambda}_{\ell m}(a)=\lambda_{\ell m}(ia)$ (cf. 21.6.4 and 21.7.5 in \cite{AbramowitzStegun64}; note $\lambda_{\ell m}(ia)\in\mathbb{R}$). Therefore, at least in this simple example it is verified that eigenvalues $\lambda $ (eigenfunctions $\psi$) of the operator $L_{\cal S}=2L[-ia]$~\footnote{$L[-ia]$ assumes implicitly a H\'aji\v {c}ek{} form $\Omega = \sin^2 \theta d\varphi$.} can be actually recovered by solving for the ``rotated'' $a\to ia$ self-adjoint operator $L[a]$, and then inverting the rotation $a\to \frac{1}{i}a=-ia$ in the resulting eigenvalues $\bar{\lambda}$ (eigenfunctions $\bar{\psi}$). \subsection{Analyticity in the ``fine structure constant''} The fact that the MOTS-stability operator can be obtained from the QCP Hamiltonian as an analytic continuation of $L[\sqrt{\alpha}]$ ($\sqrt{\alpha}=1\to\sqrt{\alpha}=-i$), together with the discussion of the previous explicit example, raise the following question: {\em can we recover the MOTS-spectrum ($\alpha=-1$) as an analytic extension of the QCP spectrum ($\alpha=1$) self-adjoint problem?} This question dwells naturally in the perturbation theory of linear operators (where $L[\sqrt{\alpha}]$ defines a self-adjoint holomorphic family of type (A) \cite{Kato80}), but giving a fully general answer defines a difficult problem. A given eigenvalue $\lambda(\sqrt{\alpha})$ can be analytically continued along its path in the complex plane, as long as its evolution does not encounter (for the same $\sqrt{\alpha}$) another eigenvalue. But checking this is a hard task even in the explicit example above. On the other hand, our particular setting is free of two potential threats for the analyticity discussion, namely boundaries and function pathologies: {\em i)} ${\cal S}$ has no boundaries (is closed), and {\em ii)} the functions in $L[\sqrt{\alpha}]$, being induced from the ambient geometry, can be taken as regular as needed. As a third point {\em iii)}, potential topological issues associated to the underlying $U(1)$ or $ \mathbb{R}^+$-fibre bundle are absent since such bundle is trivial (we are excluding here the possibility of a non-trivial NUT charge). Supported by these points and in the assumption that the example in \ref{s:explicitexample} contains all the relevant qualitative elements, we propose the following: \medskip \noindent {\bf Analyticity Conjecture.} {\em Given an orientable closed surface ${\cal S}$ and the one-parameter family of operators $L[\sqrt{\alpha}]$ defined in (\ref{e:L[alpha]}), the MOTS-spectrum ($\sqrt{\alpha} = -i$) can be recovered as an ``analytic continuation'' of the QCP spectrum ($\sqrt{\alpha} = 1$). } \medskip We present this conjecture as an open problem. In case the conjecture proves to be valid~\footnote{Even without a continuation argument, the prescription $\protect \sqrt {\alpha }\to -i\protect \sqrt {\alpha }$ to recover MOTS spectra from QCPs could still hold.}, the MOTS-stability spectrum problem would be ``essentially'' reduced to that of the self-adjoint problem of the stationary non-relativistic QCP. \subsection{Ground state of the charged particle} As a first application, we consider a transfer in the ``inverse'' sense, by using a MOTS result to calculate the ground state energy $E_o$ of QCPs. In \cite{AndMarSim07} a variational Rayleigh-Ritz-like expression for $\lambda_o$ is presented. This remarkable result does not follow from the Rayleigh-Ritz characterization, since $L_{\cal S}$ is generically not selfadjoint. The expression for $\lambda_o$ is rather obtained by starting from a min-max characterization by Donsker and Varadhan \cite{DonVar75}, valid for real not necessarily selfadjoint operators. If the conjecture above proves true, the ``rotation'' $\sqrt{\alpha}\to -i\sqrt{\alpha}$ in the $\lambda_o$ of \cite{AndMarSim07} results in a QCP $E_o$ \begin{eqnarray} \label{e:E_0_gaugeinvariant} E_o = \underset{\psi>0}{\mathrm{inf}} \int_{\cal S} \left( |D\psi|^2 + \left(e\phi + V +e^2|D\omega_\psi + z|^2\right) \psi^2 \right) \eta_{\cal S} \end{eqnarray} where $A_a = z_a + D_a \zeta$ (with $D_a z^a=0$), $\int_{\cal S} \psi^2 \eta_{\cal S} = 1$ and $\omega_\psi$ satisfies, for a given $\psi>0$, the constraint equation \begin{eqnarray} \label{e:E_0_constraint} -\Delta \omega_\psi -\frac{2}{\psi} D_a \psi D^a\omega_\psi = \frac{2}{\psi} z^a D_a\psi \ \ . \end{eqnarray} This expression for $E_o$, ``blindly'' transported from the MOTS result, has two virtues as compared with the straightforward evaluation of the Rayleigh-Ritz $E_o= \underset{||\psi||=1}{\mathrm{inf}} \int_{\cal S} \psi^* \hat{H} \psi\eta_{\cal S}$: i) it is explicitly gauge-invariant, since the term $D_aA^a(=\Delta \zeta)$ is absent; and ii) the paramagnetic term is recast as a diamagnetic one, something of potential interest in numerical strategies. Then, a crucial point is that the ``blind'' expression (\ref{e:E_0_gaugeinvariant}) for QCPs can actually be proved: starting from the (now valid) Rayleigh-Ritz expression and adapting the steps in \cite{AndMarSim07} to the MOTS/QCPs analogy, expression (\ref{e:E_0_gaugeinvariant}) follows. This is remarkable since, although it is known \cite{DonVar75} that Rayleigh-Ritz and Donsker-Varadhan expressions coincide when they both apply (namely, selfadjoint real operators), they cannot be generically reduced to one another (in particular in our setting $A_a\neq 0$). Therefore, the fact that (\ref{e:E_0_gaugeinvariant}) still holds in the fully general case offers a first non-trivial test of the conjecture. \subsection{Semi-classical approach to the MOTS spectrum} As a second example, the effective reduction to a selfadjoint problem would open a particular avenue to the use of approximate tools in the MOTS spectral problem, by transferring the semi-classical tools for QCPs \footnote{But note that ``non-selfadjoint'' semi-classical tools also exist.}. More specifically, starting from the MOTS-stability operator, we can in a first step consider the analogous QCP Hamiltonian $\hat{H}[\sqrt{\alpha}\in \mathbb{R}]$, and in a second step its corresponding classical Hamiltonian function \begin{eqnarray} \label{e:classical_Ham} H_{\mathrm{cl}}[\sqrt{\alpha}](x,p) = (p - \sqrt{\alpha}\Omega)^2 + \frac{1}{2}R_{\cal S} - G(k,\ell) \ , \end{eqnarray} defined on the cotangent bundle $T^*{\cal S}$ by reverting the ``quantization rule'', $p_i \to -i D_i$. Approximate eigenvalues and eigenfunctions for the MOTS problem could then obtained, assuming the validity of the conjecture, by applying semi-classical tools on $H_{\mathrm{cl}}[\sqrt{\alpha}]$ and then evaluating the explicit result on $\sqrt{\alpha}\to -i$. Whereas WKB techniques could be appropriate in separable problems, generic cases (notably, $\Omega_a\neq 0$) would need to resort to the rich semi-classical tools developed in the setting of quantum chaos studies (e.g. \cite{Berry83,Nonne10}). \medskip \smallskip\noindent\emph{Acknowledgments.~} \noindent I thank A. Afriat, C. Aldana, V. Aldaya, L. Andersson, M. Ansorg, A. Ashtekar, C. Barcel\'o, J. Bi\v{c}\'ak, M.E. Gabach-Cl\'ement, Q. Hummel, A. Marquina, M. Mars, G. Mena-Marug\'an, J.P. Nicolas, I. R\'acz, M. Reiris, L. Rezzolla, M. S\'anchez, J.M.M. Senovilla, W. Simon, S. Nonnenmacher, J.D. Urbina and J.A. Valiente-Kroon. I fondly thank E. Alcal\'a, whose inspiring images made this research possible.
\section{Introduction} The theories we are usually dealing with are Newtonian in the sense that the Lagrangian function depends on the first time derivatives only. There is, however, an important exception. It can happen that we are interested only in some selected degrees of freedom. By eliminating the remaining degrees one obtains what is called an effective theory. The elimination of a degree of freedom results in increasing the order of dynamical equations for remaining variables. Therefore, effective theories are described by Lagrangians containing higher order time derivatives \cite{b1}. Originally, these theories were proposed as a method for dealing with ultraviolet divergences \cite{b2}; this idea appeared to be quite successful in the case of gravity: the Einstein action supplied by the terms containing higher powers of curvature leads to a renormalizable theory \cite{b3}. Other examples of higher derivatives theories include the theory of the radiation reaction \cite{b4,b4a}, the field theory on noncommutative spacetime \cite{b5,b5a}, anyons \cite{b6,b6a} or string theories with the extrinsic curvature \cite{b7}. \par Of course, the appearance of terms with higher time derivatives leads to some problems. One of them is that the energy does not need to be bounded from below. To achieve a deeper insight into these problems and, possibly, to find a solution it is instructive to consider a quite simple, however nontrivial, higher derivatives theory. For example, it was shown in Ref. \cite{b8} (see also \cite{b8a}) that the problem of the energy can be avoided (on the quantum level) in the case of the celebrated Pais-Uhlenbeck (PU) oscillator \cite{b9}. This model has been attracting considerable interest throughout the years (for the last few years, see, e.g., \cite{b8,b8a},\cite{b10a}-\cite{ b11}). Recently, it has been shown (see, \cite{b11}) that the properties of the PU oscillator, rather surprisingly, for some special values of frequencies change drastically and are related to nonrelativistic conformal symmetries. Namely, if the frequencies of oscillations are {\it odd} multiplicities of a basic one, i.e., they form an arithmetic sequences $\omega_k=(2k-1)\omega,\quad \omega\neq 0$, for $k=1,\ldots,n$, then the maximal group of Noether symmetries of the PU Lagrangian is the $l$-conformal Newton-Hooke group with $l=\frac{2n-1}{2}$ (for more informations about these groups see, e.g., \cite{b12a}-\cite{b12d} and the references therein). Otherwise, the symmetry group is simpler (there are no counterparts of dilatation and conformal generators (see, the algebra (\ref{e5})). \par Much attention has been also paid to Hamiltonian formulations of the PU oscillator. There exists a few approaches to Hamiltonian formalism of the PU model: decomposition into the set of the independent harmonic oscillators proposed by Pais and Uhlenbeck in their original paper \cite{b9}, Ostrogradski approach based on the Ostrogradski method \cite{b13} of constructing Hamiltonian formalism for theories with higher time derivatives and the last one, applicable in the case of odd frequencies (mentioned above), which exhibits the $l$-conformal Newton-Hooke group structure of the model. Consequently, there arises a natural question about the relations between them as well as the realization of the symmetry on the Hamiltonian level? The aim of this work is to give the answer to this question. \par The paper is organized as follows. After recalling, in Section 2, some informations concerning symmetry of the PU model on the Lagrangian level, we start with the harmonic decoupled approach. We find, on the Hamiltonian level, the form of generators (for both generic and odd frequencies) and we show that they, indeed, form the algebra which is central extension the one appearing on the Lagrangian level. Section 4 is devoted to the study of the relation between the above approach and the Ostrogradski one. Namely, we construct the canonical transformation which relates the Ostrogradski Hamiltonian to the one describing the decouple harmonic oscillator. This transformation enables us to find the remaining symmetry generators in terms of Ostrogradski variables. The next section is devoted to the case of odd frequencies where the additional natural approach can be constructed. In this framework the Hamiltonian is the sum of the one for the free higher derivatives theory and the conformal generator. We derive a canonical transformation which relates this new Hamiltonian to the one for the PU oscillator with odd frequencies. Moreover, we apply the method (see, \cite{b14}) of constructing integrals of motion for the systems with symmetry to find all symmetry generators. Next, by direct calculations we show that they are related by the, above mentioned, canonical transformation to the ones of the PU model described in terms decoupled oscillators. We also express symmetry generators in terms of their counterparts in the free theory. \par In concluding Section 6, we summarize our results and discuss possible further developments. Finally, Appendix constitutes technical support for the mains results. We derive there some relations and identities which are crucial for our work. \section{PU oscillator and its symmetry} Let us consider the three-dimensional PU oscillator, i.e., the system which is described by the following Lagrangian \cite{b9} \be \label{e1} L=-\frac{1}{2}\vec x \prod_{k=1}^{n}\left(\frac{d^2}{dt^2}+\omega_k^2\right)\vec x, \ee where $0<\omega_1<\omega_2<\ldots<\omega_n$ and $n=1,2,\ldots$. Lagrangian (\ref{e1}) implies the following equation of motion \be \label{e2} \prod_{k=1}^n\left(\frac{d^2}{dt^2}+\omega_k^2\right)\vec x=0, \ee which possesses the general solution of the form \be \label{e3} \vec x(t)=\sum_{k=1}^{n}(\vec \alpha_k\cos\omega_kt+\vec \beta_k\sin\omega_kt), \ee where $\vec \alpha's$ and $\vec \beta's$ are some arbitrary constants. \par As it has been mentioned in the Introduction the structure of the maximal symmetry group of Lagrangian (\ref{e1}) depends on the values of $\omega's$. If the frequencies of oscillation are { odd}, i.e., they form an arithmetic sequence $\omega_k=(2k-1)\omega,\quad \omega\neq 0$, $k=1,\ldots,n$, then the maximal group of Noether symmetries of the system (\ref{e1}) is the $l$-conformal Newton-Hooke group, with $l=\frac{2n-1}{2}$. It is the group which Lie algebra is spanned by $H,D,K,J^{\alpha\beta}$ and $C_p^\alpha$, $\alpha,\beta=1,2,3$, $p=0,1,\ldots,2n-1$, satisfying the following commutation rules \be \begin{split} \label{e4} [H,D]&=H-2\omega^2 K,\quad [H,K]=2D,\quad [D,K]=K,\\ [D,\vec C_p]&=(p-\frac{2n-1}{2})\vec C_p,\quad [K,\vec C_p]=(p-2n+1)\vec C_{p+1},\\ [H,\vec C_p]&=p\vec C_{p-1}+(p-2n+1)\omega^2\vec C_{p+1},\\ [J^{\alpha\beta},J^{\gamma\delta}]&=\delta^{\alpha\delta}J^{\gamma\beta}+\delta^{\alpha\gamma}J^{\beta\delta}+\delta^{\beta\gamma}J^{\delta\alpha}+\delta^{\beta\alpha}J^{\alpha\gamma},\\ [J^{\alpha\beta},C_p^\gamma]&=\delta^{\alpha\gamma}C_p^\beta-\delta^{\beta\gamma}C_p^\alpha. \end{split} \ee Although this algebra is isomorphic to the $l$-conformal Galilei one (the latter can be obtained by a linear change of the basis $H\rightarrow H-\omega^2K$, see \cite{b12a,b12b,b12d} and \cite{b15a}-\cite{b15f} for more recent developments of this algebra) the use of the basis (\ref{e4}) implies the change of the Hamiltonian which alters the dynamics. \par In the case of {\it generic} frequencies the maximal symmetry group is simpler. Its Lie algebra consists of $H,J^{\alpha\beta}$ and $\vec C^{\pm}_k$, $k=1,\ldots,n$. The action of $J^{\alpha\beta}$ remains unchanged and only commutations rules between $H$ and $\vec C's$ must be modified \be \label{e5} \begin{split} [H,\vec C_k^+]&=-\omega_k\vec C_k^-,\\ [H,\vec C_k^-]&=\omega_k\vec C_k^+. \end{split} \ee \par Both symmetry algebras posses central extension: \be \label{e6} [C_p^\alpha,C_q^\beta]=(-1)^pp!q!\delta_{\alpha\beta}\delta_{2n-1,p+q}, \ee in the odd case and \be \label{e7} [C_k^{+\alpha}C_{j}^{-\beta}]=\frac{\omega_k}{\rho_k}\delta_{kj}\delta^{\alpha\beta}, \ee in the generic case; which will turn out to be necessarily (see the next section) to construct the symmetry algebra on the Hamiltonian level. \section{Decoupled oscillators approach} An approach to the Hamiltonian formalism of the PU model was proposed in Ref. \cite{b9} where it was demonstrated that the Hamiltonian of the PU oscillator (in dimension one) turns into the sum of the harmonic Hamiltonians with alternating sign. To show this we follow the reasoning of Ref. \cite{b9} and introduce new variables \be \label{e8} \vec x_k=\Pi_k\vec x, \quad k=1,\ldots,n; \ee where $\Pi_k$ is the projective operator: \be \label{e9} \Pi_k=\sqrt{|\rho_k|}\prod_{\substack{i=1 \\ i\neq k}}^n\left(\frac{d^2}{dt^2}+\omega_i^2\right) , \ee and \be \label{r0} \rho_k=\frac{1}{\prod\limits_{\substack{i=1 \\ i\neq k}}^n(\omega_i^2-\omega_k^2)},\quad k=1,2,\ldots,n. \ee Note that $\rho_k$ are alternating in sign. Then one finds \be \label{e10} \vec x=\sum_{k=1}^n(-1)^{k-1}\sqrt{|\rho_k|}\vec x_k, \ee as well as \be \label{e11} L=-\frac{1}{2}\sum_{k=1}^{n}(-1)^{k-1}\vec x_k\left(\frac{d^2}{dt^2}+\omega_k^2\right)\vec x_k=\frac{1}{2}\sum_{k=1}^{n}(-1)^{k-1}(\dot{\vec{x_k}}-\omega_k^2\vec x_k^2)+t.d. \ee The corresponding Hamiltonian reads \be \label{e12} H=\frac{1}{2}\sum_{k=1}^{n}(-1)^{k-1}(\vec p_k^2+\omega_k^2\vec x_k^2), \ee while the canonical equations of motion are of the form \be \label{e13} \dot {\vec x}_k=(-1)^{k-1}\vec p_k,\quad \dot{\vec p}_k=(-1)^k\omega_k^2 \vec x_k. \ee Taking into account the form of the general solution (\ref{e3}) we see that the dynamics of the new canonical variables is given by \be \begin{split} \label{e14} \vec x_k=\frac{(-1)^{k-1}}{\sqrt{|\rho_k|}}(\vec \alpha_k\cos(\omega_kt)+\vec \beta_k\sin(\omega_kt)),\\ \vec p_k=\frac{\omega_k}{\sqrt{|\rho_k|}}(\vec \beta_k\cos(\omega_kt)-\vec \alpha_k\sin(\omega_kt)). \end{split} \ee Therefore, we have a correspondence between the set of solutions of the Lagrange equation (\ref{e2}) and the set of solutions of the canonical equations (\ref{e13}). Consequently, we can translate the action of the group symmetry from the Lagrangian level to the Hamiltonian one and find all the symmetry generators in terms of oscillator canonical variables. We will show that the generators, obtained in this way, form the algebra which is the central extension of the symmetry algebra on the Lagrangian level. \par In the generic case it is very easy to find the form of the remaining (the Hamiltonian is given by (\ref{e12})) symmetry generators on the Hamiltonian level. First, let us note that the infinitesimal action of $\vec\mu _k\vec {C}^{+}_k$ and $\vec\nu _k\vec {C}^{-}_k$, $k=1,\ldots,n$, on the Lagrangian level, takes the form \be \label{e15} \begin{split} \vec x'(t)=\vec x(t)+\sum_{k=1}^n(\vec \mu_k\cos \omega_k t+\vec\nu_k \sin\omega_k t).\ \end{split} \ee Acting with $\Pi_k$ and applying Eq. (\ref{e13}) we find the infinitesimal action of $\vec {C}^{\pm}_k$ on the phase space; by virtue of \be \label{e16} \delta(\cdot)=\epsilon \{\cdot, \textrm{Generator}\}, \ee we obtain the following generators: \be \begin{split} \label{e17} \vec C_k^+&=\frac{(-1)^{k-1}}{\sqrt{|\rho_k|}}\cos(\omega_k t)\vec p_k+\frac{\omega_k}{\sqrt{|\rho_k|}}\sin(\omega_k t)\vec x_k,\\ \vec C_k^-&=\frac{(-1)^{k-1}}{\sqrt{|\rho_k|}}\sin(\omega_k t)\vec p_k-\frac{\omega_k}{\sqrt{|\rho_k|}}\cos(\omega_k t)\vec x_k, \end{split} \ee which commute to the central charge -- according to (\ref{e7}). Similarly, the angular momentum generators read \be \label{e18} J^{\alpha\beta}=\sum_{k=1}^n( x^\alpha_k p_k^\beta-p_k^\alpha x_k^\beta ). \ee Consequently, we obtain the centrally extended algebra (\ref{e5}). \subsection{Odd frequencies} In the odd case the symmetry group is reacher and, therefore, this case is much more interesting. We assume now that the frequencies form the arithmetic sequence, i.e., $\omega_k=(2k-1)\omega$, $k=1,\ldots,n$. In this case the main point is that the numbers $\rho_k$ can be explicitly computed; the final result reads \be \label{r1} \rho_k=\frac{(-1)^{k-1}(2k-1)}{(4\omega^2)^{n-1}(n-k)!(n+k-1)!}, \quad k=1,\ldots,n. \ee Consequently, one has useful relations \be \label{r2} \frac{|\rho_k|}{|\rho_{k+1}|}=\frac{(2k-1)(n+k)}{(2k+1)(n-k)},\quad k=1,\ldots,n-1. \ee Next, let us note that the following Fourier expansion holds (see, Appendix) \be \label{e20} \sin^p \wt\cos^{2n-1-p}\wt = \left\{ \begin{array}{c} \sum\limits_{k=1}^m\gamma^+_{kp}\cos(2k-1)\wt, \quad p \textrm{ - even};\\ \sum\limits_{k=1}^m\gamma^-_{kp}\sin(2k-1)\wt, \quad p \textrm{ - odd}; \end{array} \right. \ee where $\gamma^\pm_{kp}$ can be expressed in terms of sum of products of binomial coefficients; however, their explicit form is not very useful; for our purposes some properties of $\gamma^\pm_{kp}$ (see, (\ref{a2})-(\ref{a6})) will turn out to be more fruitful. Now, using Eq. (\ref{e20}) we can rewrite the infinitesimal action (\ref{e15}), in the case of odd frequencies, in the equivalent form \be \label{e19} \vec x'(t)=\vec x(t)+\frac{1}{\omega^p}\vec \epsilon_p\sin^p\wt\cos^{2n-1-p}\wt, \ee which gives suitable family of the generators $\vec C_p$, $p=0,1,2,\ldots,2n-1$ on the Lagrangian level, i.e., satisfying commutation rules of the $l$-conformal Newton-Hooke algebra (cf., \cite{b11}) \par In order to find the action of $\vec C_p$ in the Hamiltonian formalism, we use Eqs. (\ref{e20}) together with (\ref{e8}) and (\ref{e13}), which yields \be \label{e21} \vec x'_k=\vec x_k+\frac{(-1)^{k-1}\vec \epsilon_p}{\omega^p\sqrt{|\rho_k|}}\left\{ \begin{array}{c} \gamma^+_{kp}\cos(2k-1)\wt,\quad p \textrm{ - even};\\ \gamma^-_{kp}\sin(2k-1)\wt,\quad p \textrm{ - odd}; \end{array} \right. \ee \be \label{e22} \vec p'_k=\vec p_k+\frac{(2k-1)\omega\vec \epsilon_p}{\omega^p\sqrt{|\rho_k|}}\left\{ \begin{array}{c} -\gamma^+_{kp}\sin(2k-1)\wt,\quad p \textrm{ - even};\\ \gamma^-_{kp}\cos(2k-1)\wt,\quad p \textrm{ - odd}. \end{array} \right. \ee Using Eq. (\ref{e16}) we derive the explicit expression for the generators $\vec C_p$ in terms of the canonical variables \be \label{e23} \vec C_p= \sum\limits_{k=1}^{n}\frac{\gamma_{kp}^+}{\omega^p\sqrt{|\rho_k|}}\left( (-1)^{k-1}\cos((2k-1)\wt)\vec p_k+(2k-1)\omega\sin((2k-1)\wt)\vec x_k\right),\\ \ee for $ p$ even, and \be \label{e23a} \vec C_p= \sum\limits_{k=1}^{n}\frac{\gamma_{kp}^-}{\omega^p\sqrt{|\rho_k|}}\left( (-1)^{k-1}\sin((2k-1)\wt)\vec p_k-(2k-1)\omega\cos((2k-1)\wt)\vec x_k\right), \ee for $p$ odd. Eqs. (\ref{e23}) and (\ref{e23a}) can be inverted to yield $\vec x_k$ and $\vec p_k$ in terms of the generators $\vec C_p$ \be \label{e24} \begin{split} \vec p_k&=(-1)^{k-1}\sqrt{|\rho_k |}\cos((2k-1)\wt)\sideset{}{''}\sum_{p=0}^{2n-1}\beta^+_{pk}\omega^p\vec C_p\\ &+(-1)^{k-1}\sin((2k-1)\wt)\sideset{}{'}\sum_{p=0}^{2n-1}\beta^-_{pk}\omega^p\vec C_p , \end{split} \ee \be \label{e25} \begin{split} \vec x_k&=\frac{\sqrt{|\rho_k |}}{(2k-1)\omega}\sin((2k-1)\wt)\sideset{}{''}\sum_{p=0}^{2n-1}\beta^+_{pk}\omega^p\vec C_p\\ &-\frac{\sqrt{|\rho_k |}}{(2k-1)\omega}\cos((2k-1)\wt)\sideset{}{'}\sum_{p=0}^{2n-1}\beta^-_{pk}\omega^p\vec C_p , \end{split} \ee where $\beta^+,\beta^-$ are the inverse matrices to $\gamma^+,\gamma^-$ while one and two primes $',''$ denote the sum over odd and even indices, respectively\footnote{We will use this convention throughout the article.}. \par Next, we find the action of the dilatation generator. To this end let us recall (cf., \cite{b11}) that the infinitesimal action of dilatation on coordinates is of the form \be \label{e26} \vec x'(t)=\vec x(t)-\frac{\epsilon}{2\omega}\left((2n-1)\omega\cos(2\wt)\vec x(t)-\sin(2\wt)\dot{\vec x}(t)\right). \ee Substituting (\ref{e10}) and acting with the projectors $\Pi_k$ we obtain, due to (\ref{e8}) and (\ref{e13}), the infinitesimal dilatation transformation on the phase space \be \label{e27} \begin{split} \vec x_k'&=\vec x_k+\frac{\epsilon}{2\sqrt{|\rho_k|}}\cos(2\wt)\left(\sqrt{|\rho_{k-1}|}(n-k+1)\vec x_{k-1}+\sqrt{|\rho_{k+1}|}(n+k)\vec x_{k+1}\right) \\ &+\frac{\epsilon(-1)^{k}}{2\omega\sqrt{|\rho_k|}}\sin(2\wt)\left(\frac{\sqrt{|\rho_{k-1}|}}{2k-3}(n-k+1)\vec p_{k-1}- \frac{\sqrt{|\rho_{k+1}|}}{2k+1}(n+k)\right)\vec p_{k+1},\\ \vec x_1'&=\vec x_1-\frac{\epsilon}{2\sqrt{|\rho_1|}}\left(\sqrt{|\rho_{2}|}(n+1)\cos(2\wt)\vec x_{2}+\sin(2\wt)\sqrt{|\rho_{2}|}\frac{(n+1)}{3\omega}\vec p_{2} \right.\\ &\left. -n\cos(2wt)\sqrt{|\rho_1|}\vec x_1+\frac n\omega\sin(2\wt)\sqrt{|\rho_1|} \vec p_1\right), \end{split} \ee \be \label{e28} \begin{split} \vec p_k'&=\vec p_k-\frac{\epsilon(2k-1)}{2\sqrt{|\rho_k|}}\cos(2\wt)\left(\frac{\sqrt{|\rho_{k-1}|}}{2k-3}(n-k+1)\vec p_{k-1}+ \frac{\sqrt{|\rho_{k+1}|}}{2k+1}(n+k)\vec p_{k+1}\right)\\ &+\frac{\epsilon\omega(-1)^{k}(2k-1)}{2\sqrt{|\rho_k|}}\sin(2\wt)\left(\sqrt{|\rho_{k-1}|}(n-k+1)\vec x_{k-1}-\sqrt{|\rho_{k+1}|}(n+k)\vec x_{k+1}\right) ,\\ \vec p_1'&=\vec p_1-\frac{\epsilon}{2\sqrt{|\rho_1|}}\cos(2\wt)\left(-n\sqrt{|\rho_{1}|}\vec p_{1}+ \frac{\sqrt{|\rho_{2}|}}{3}(n+1)\vec p_{2}\right)\\ &+\frac{\epsilon\omega}{2\sqrt{|\rho_1|}}\sin(2\wt)\left(n\sqrt{|\rho_{1}|}\vec x_{1}+\sqrt{|\rho_{2}|}(n+1)\vec x_{2}\right) ,\\ \end{split} \ee where $k>1$ and, by definition, we put $\vec x_{n+1}=\vec p_{n+1}=0$. One can check, using Eq. (\ref{r2}), that (\ref{e27}) and (\ref{e28}) define the infinitesimal canonical transformation generated (according to (\ref{e16})) by \be \label{e29} D=\frac{-1}{2\omega}\left(\omega A\cos(2\wt)+B\sin(2\wt)\right), \ee where \be \label{e30} \begin{split} A=&-\sum_{k=1}^n\left(\sqrt{\left|\frac{\rho_{k-1}}{\rho_{k}}\right|}(n-k+1)\vec x_{k-1}+\sqrt{\left|\frac{\rho_{k+1}}{\rho_k}\right|}(n+k)\vec x_{k+1}\right)\vec p_k+ n\vec x_1\vec p_1,\\ B=&-\sum_{k=1}^n(-1)^{k}\frac{n-k+1}{2k-3}\sqrt{\left| \frac{\rho_{k-1}}{\rho_{k}}\right| }(\vec p_k\vec p_{k-1}-(2k-1)(2k-3)\omega^2\vec x_k\vec x_{k-1})\\ &+\frac{1}{2}n(\omega^2\vec x_1^2-\vec p_1^2), \end{split} \ee and, by definition, $\vec x_0=\vec p_0=0$. The meaning of the components $A$ and $B$ will become more clear in Section 5 (see, (\ref{e50})). \par Similar calculations can be done for the conformal generator $K$. Namely, the infinitesimal conformal transformation, on the Lagrangian level, reads \be \label{e31} \vec x'(t)=\vec x(t)-\frac{\epsilon}{2\omega^2}\left((2n-1)\omega\sin(2\wt)\vec x(t)+(\cos(2\wt)-1)\dot{\vec x}(t)\right). \ee \par Substituting $\vec x$ and acting with the projector $\Pi_k$ we obtain the infinitesimal conformal transformation on the phase space and consequently (due to (\ref{e16})) the explicit form of the generator $K$ \be \label{e32} K=\frac{1}{2\omega^2}\left(B\cos(2\wt)-\omega A\sin(2\wt)+H\right). \ee Finally, the angular momentum takes the same form as in the generic case \be \label{e33} J^{\alpha\beta}=\sum_{k=1}^n( x^\alpha_k p_k^\beta-p_k^\alpha x_k^\beta ). \ee \par It remains to verify that obtained generators, indeed, yield integrals of motion and define the centrally extended $l$-conformal Newton-Hooke algebra. To this end we need a few identities which are proven in the Appendix. First, we compute the commutators of $\vec C$'s and check that they give the proper central extension. The only nontrivial case is $[C_p^\alpha, C_q^\beta]$ with $p$ even and $q$ odd (or conversely). We have \be \label{e34} \begin{split} [ C_p^\alpha, C_q^\beta]&=\frac{\omega(2\omega)^{2(n-1)}\delta^{\alpha\beta}}{\omega^{p+q}} \sum_{k=1}^n(-1)^{k-1}(n-k)!(n+k-1)!\gamma^+_{kp}\gamma^-_{kq}\\ & =\frac{p!(2n-1-p)!\omega^{2n-1}\delta^{\alpha\beta}}{\omega^{p+q}} \sum_{k=1}^n(-1)^{k-1}\beta^+_{pk}\gamma^-_{kq}\\ &=\frac{p!(2n-1-p)!\omega^{2n-1}\delta^{\alpha\beta}}{\omega^{p+q}} \sum_{k=1}^n\beta^-_{2n-1-p,k}\gamma^-_{kq}=\\ &=\frac{p!(2n-1-p)!\omega^{2n-1}\delta^{\alpha\beta}}{\omega^{p+q}} \delta_{2n-1,p+q}=p!q!\delta_{\alpha\beta}\delta_{2n-1,p+q}, \end{split} \ee where we use consecutively Eqs. (\ref{e23}), (\ref{e23a}), (\ref{r1}), (\ref{a3}) and (\ref{a2}). For $p$ odd and $q$ even we obtain the same result except the extra minus sign. Consequently, we obtain the central extension (\ref{e6}). In order to find the remaining commutators let us note that \be \label{e35} \begin{split} [A,B]&=-2H,\\ [B,H]&=2\omega^2A,\\ [A,H]&=-2B. \end{split} \ee The proof of the above relations is straightforward although tedious and involve the use of (\ref{r2}). Now, by virtue of Eq. (\ref{e35}), it is easy to check that the generators $H,D,K$ satisfy the first line of equations (\ref{e4}). \par Now, we find the adjoint action of $H,D,K,J^{\alpha\beta}$ on $\vec C_p$. Since the calculations are rather wearisome and lengthy we sketch only the main points. To show that $ [H,\vec C_p] $ gives proper rule we use the identity (\ref{a4}). The case $[D,\vec C_p]$ is more involved; however, using repeatedly Eqs. (\ref{r2}) and (\ref{a5}) we arrive at the desired result. Similarly to obtain $[H,\vec C_k]$, first, we use Eq. (\ref{r2}) and then Eq. (\ref{a6}). Finally, it is easy to compute the commutators involving angular momentum. \par Having all the commutation rules and (\ref{a4}) it is not hard to check that the obtained generators are constants of motion. This concludes the proof that, on the Hamiltonian level, they are symmetry generators and form the centrally extended $l$-conformal Newton-Hooke algebra. \section{Ostrogradski approach} Since the PU oscillator is an example of higher derivatives theory, it is natural to use the Hamiltonian formalism proposed by Ostrogradski \cite{b13}. To this end let us expand Lagrangian (\ref{e1}) in the sum of higher derivatives terms (here, $\vec Q=\vec x$) \be \label{e36} L=-\frac{1}{2}\vec Q \prod_{k=1}^{n}\left(\frac{d^2}{dt^2}+\omega_k^2\right)\vec Q=\frac{1}{2}\sum_{k=0}^n(-1)^{k-1}\sigma_k(\vec Q^{(k)})^2 , \ee where \be \label{e37} \sigma_k=\sum_{i_1<\ldots<i_{n-k}}\omega_{i_1}^2\cdots \omega_{i_{n-k}}^2,\quad k=0,\ldots,n; \quad \sigma_n=1. \ee It can be shown (by standard reasoning) that the following identities hold \begin{align} \label{r3}&\sum_{k=1}^n\rho_k\omega_k^{2m}=0, \qquad m=0,\ldots ,n-2, \\ \label{r4}&\sum_{k=1}^n\rho_k\omega_k^{2(n-1)}=(-1)^{n+1},\\ \label{r5}&\sum_{m=0}^n\sigma_m(-1)^m\sum_{k=1}^n\rho_k\omega_k^{2(r-n+m-1)}=0, \quad r\geq n, \end{align} where $\rho_k$ is given by Eq. (\ref{r0}). Now, we introduce the Ostrogradski variables \be \label{e38} \begin{split} \vec Q_k&=\vec Q^{(k-1)}, \\ \vec P_k&=\sum_{j=0}^{n-k}\left(-\frac{d}{dt}\right)^j\frac{\partial L}{\partial \vec Q^{(k+j)}}=(-1)^{k-1}\sum_{j=k}^n\sigma_j\vec Q^{(2j-k)}, \end{split} \ee for $k=1,\ldots,n$. Then the Ostrogradski Hamiltonian takes the form \be \label{e39} H=\frac{(-1)^{n-1}}{2}\vec P_n^2+\sum_{k=2}^n\vec P_{k-1}\vec Q_k-\frac 1 2 \sum_{k=1}^n(-1)^k\sigma_{k-1}\vec Q_{k}^2. \ee By virtue of Eqs. (\ref{e13}) and (\ref{e38}), for $k=1,\ldots,n$, we find \be \begin{split} \label{e40} \vec Q_k&=(-1)^{\frac{k-1}{2}}\sum_{j=1}^n\sqrt{|\rho_j|}(-1)^{j-1}\omega_j^{k-1}\vec x_j,\quad k-\textrm {odd};\\ \vec Q_k&=(-1)^{\frac{k}{2}-1}\sum_{j=1}^n\sqrt{|\rho_j|}\omega_j^{k-2}\vec p_j,\quad k- \textrm {even};\\ \end{split} \ee and \be \label{e41} \begin{split} \vec P_k&=(-1)^{\frac k2-1}\sum_{i=1}^n(-1)^{i-1}\sqrt{|\rho_i|}\left(\sum_{j=k}^n\sigma_j(-1)^j\omega_i^{2j-k}\right)\vec x_i, \quad k-\textrm{ even};\\ \vec P_k&=(-1)^{\frac {k-3}{2}}\sum_{i=1}^n\sqrt{|\rho_i|}\left(\sum_{j=k}^n\sigma_j(-1)^j\omega_i^{2j-k-1}\right)\vec p_i, \quad k- \textrm{ odd}. \end{split} \ee One can show that Eqs. (\ref{e40}) and (\ref{e41}) define a canonical transformation; to compute the Poisson brackets $\{\vec Q_k,\vec Q_j\}$ and $\{\vec Q_k $, $\vec P_j\}$ we use (\ref{r3}) and (\ref{r3})-(\ref{r5}), respectively; computing $\{\vec P_k,\vec P_j\}$ is the most complicated one and involves considering two cases $k-j\lessgtr 1$ as well as applying Eqs. (\ref{r3}) and (\ref{r5}). \par Next, let us note that the inverse transformation is of the form \be \begin{split} \label{e42} \vec x_i&=\sideset{}{'}\sum_{k=1}^{n}(-1)^{\frac{k-3}{2}}\sum_{j=k}^n\sigma_j(-1)^j\omega_i^{2j-k-1}\sqrt{|\rho_i|}\vec Q_k+\sideset{}{''}\sum_{k=1}^{n}(-1)^{\frac{k}{2}}\sqrt{|\rho_i|}\omega_i^{k-2}\vec P_k,\\ \vec p_i&=\sideset{}{''}\sum_{k=1}^{n}(-1)^{\frac{k}{2}+i-1}\sum_{j=k}^n\sigma_j(-1)^j\omega_i^{2j-k}\sqrt{|\rho_i|}\vec Q_k+\sideset{}{'}\sum_{k=1}^{n}(-1)^{\frac{k+1}{2}+i}\sqrt{|\rho_i|}\omega_i^{k-1}\vec P_k. \end{split} \ee \par No, we can try to find the symmetry generators (both in the odd and generic cases) in terms of the Ostrogradski variables. Of course, we expect that the Hamiltonian (\ref{e12}) should be transformed into the Ostrogradski one. Indeed, using (\ref{r3})-(\ref{r5}) repeatedly we arrive, after straightforward but rather arduous computations (considering two cases: $n$ - odd, even), at the Ostrogradski Hamiltonian (\ref{e39}). \par Similarly, applying Eqs. (\ref{r3})-(\ref{r5}), we check that the angular momentum (in both cases (\ref{e18}) and (\ref{e33})) transforms under (\ref{e42}) into Ostrogradski angular momentum \be \label{e43} J^{\alpha\beta}=\sum_{k=1}^n(Q^\alpha_kP^\beta_k-Q^\beta_k P^\alpha_k ). \ee As far as the generators $\vec C's$ are concerned (again using (\ref{r3})-(\ref{r5})) we obtain the following expressions: \be \label{e44} \begin{split} \vec C_k^-=\sum_{k=1}^n(\cos\omega_it)^{(k-1)}\vec P_k-\sum_{k=1}^n\left((-1)^{k-1}\sum_{j=k}^n\sigma_j(\cos\omega_it)^{2j-k}\right)\vec Q_k,\\ \vec C_k^-=\sum_{k=1}^n(\sin\omega_it)^{(k-1)}\vec P_k-\sum_{k=1}^n\left((-1)^{k-1}\sum_{j=k}^n\sigma_j(\sin\omega_it)^{2j-k}\right)\vec Q_k, \end{split} \ee in the case of generic frequencies, and \be \label{e45} \begin{split} \vec C_p={\frac{1}{\omega^p}}\sum_{k=1}^n\left( \vec P_k(\sin^p\wt\cos^{2n-1-p}\wt)^{(k-1)} \right.\\ \left. +(-1)^k\vec Q_k\sum_{j=k}^n\sigma_j(\sin^p\wt\cos^{2n-1-p}\wt)^{(2j-k)}\right), \end{split} \ee in the odd case; which perfectly agrees with the definitions of the Ostrogradski canonical variables (\ref{e38}) and the action of $\vec C's$ on $Q$ (Eqs. (\ref{e15}) and (\ref{e19})). Similar reasoning can be done for the remaining two generators $D$ and $K$ in the odd case. Then, they become bilinear forms in the Ostrogradski variables; however the explicit form of coefficients is difficult to simplify and not transparent thus we skip it here. \section{Algebraic approach to odd case} Since the $l$-conformal Newton-Hooke algebra is related to the $l$-conformal Galilei one by the change of Hamiltonian \be \label{e46} H=\tilde H+\omega ^2\tilde K, \ee where tilde refers to generators of the free theory (which possesses the $l$-conformal Galilei symmetry); therefore, it would be instructive to construct an alternative Hamiltonian formalism for the PU-model (in the case of {\it odd} frequencies) with the help of the one for the {\it free} higher derivatives theory. \par Denoting by $\vec q_m,\vec \pi_m$, $ {m=0},\ldots,{n-1}$ the phase space coordinates of the free theory and adapting the results of Ref. \cite{b16} to our conventions we obtain the following form of the generators of the free theory (at time $t=0$) \be \label{e47} \begin{split} \tilde H&=\frac{(-1)^{n+1}}{2}\pi_{n-1}^2-\sum_{m=1}^{n-1}\vec q_m\vec \pi_{m-1},\\ \tilde D&=\sum_{m=0}^{n-1}\left(m-\frac{2n-1}{2}\right)\vec q_m\vec\pi_m,\\ \tilde K&=(-1)^{n+1}\frac{n^2}{2}\vec q_{n-1}^2+\sum_{m=0}^{n-2}(2n-1-m)(m+1)\vec q_m\vec \pi_{m+1},\\ \tilde J^{\alpha\beta}&=\sum_{m=0}^{n-1}(q^\alpha_m\pi^\beta_m-q^\beta_m\pi^\alpha_m),\\ \tilde{\vec C}_m&=(-1)^{m+1}m!\vec \pi_m, \quad m=0,\ldots,n-1,\\ \tilde{\vec C}_{2n-1-m}&=(2n-1-m)!\vec q_m, \quad m=0,\ldots,n-1. \end{split} \ee \par Of course, the change of the algebra basis given by (\ref{e46}) induces the corresponding one for the coordinates in dual space of the algebra (denoted in the same way); consequently we define the new Hamiltonian as follows \be \begin{split} \label{e48} H&=\tilde H+\omega^2\tilde K=\frac{(-1)^{n+1}}{2}\pi_{n-1}^2-\sum_{m=1}^{n-1}\vec q_m\vec \pi_{m-1}\\ &+(-1)^{n+1}\frac{n^2\omega^2}{2}\vec q_{n-1}^2+\sum_{m=0}^{n-2}(2n-1-m)(m+1)\omega^2\vec q_m\vec \pi_{m+1}. \end{split} \ee We will show that (\ref{e48}) is indeed the PU Hamiltonian in $ \vec q_m,\vec \pi_m$ coordinates and we will find the remaining generators in terms of them. To this end let us define the following transformation \be \label{e49} \begin{split} \vec x_k&=(-1)^k\left(\sideset{}{''}\sum_{m=0}^{n-1}\frac{\omega^{-m}}{m!\sqrt{|\rho_k|}}\gamma^+_{km}\vec q_m+\sideset{}{'}\sum_{m=0}^{n-1}\frac{m!\omega^m\sqrt{|\rho_k|}}{(2k-1)\omega}\beta^+_{2n-1-m, k}\vec \pi_m\right),\\ \vec p_k&=(-1)^{k}\left(\sideset{}{'}\sum_{m=0}^{n-1}\frac{\omega^{-m}(2k-1)\omega}{m!\sqrt{|\rho_k|}}\gamma^+_{k, 2n-1-m}\vec q_m+\sideset{}{''}\sum_{m=0}^{n-1}{m!\omega^m\sqrt{|\rho_k|}}\beta^+_{m k}\vec \pi_m\right), \end{split} \ee for $k=1,\ldots,n$. Using (\ref{r2}) and (\ref{a3}) we check that (\ref{e49}) define a canonical transformation. Moreover, by applying Eqs. (\ref{r2}) and (\ref{a2})-(\ref{a4}) we check that the PU Hamiltonian (\ref{e12}) (with odd frequencies ) transforms into (\ref{e48}). The remaining generators can be also transformed. First, using (\ref{r2}), (\ref{a2}), (\ref{a3}), (\ref{a5}) and (\ref{a6}), after troublesome computations, we find that \be \label{e50} \begin{split} A&=-2\tilde D,\\ B&=-\tilde H+\omega^2\tilde K,\ \end{split} \ee and, consequently, we obtain a nice interpretation of $A$ and $B$. Using Eqs. (\ref{e50}), one checks that $H,D,K$ take the form \be \label{e51} \begin{split} H&=\tilde H+\omega^2 \tilde K,\\ D&=\tilde D \cos 2\wt+\frac{1}{2\omega}(\tilde H-\omega^2\tilde K)\sin 2\wt,\\ K&=\frac 12 (1+\cos 2\wt )\tilde K+ \frac {1}{2\omega^2} (1-\cos 2\wt )\tilde H+\frac {\sin 2\wt }{\omega } \tilde D.\\ \end{split} \ee Finally, the angular momentum reads \be \label{e52} J^{\alpha\beta}=\sum_{m=0}^{n-1}(q^\alpha_m\pi^\beta_m-q^\beta_m\pi^\alpha_m) , \ee i.e., takes the same form as the one for the free theory (according to it commutes with $H$). The generators $\vec C_k$ are obtained by plugging (\ref{e49}) into (\ref{e23}) and (\ref{e23a}), see also (\ref{e62}). \par Summarizing, we expressed all PU symmetry generators in terms of the ones for free theory (and consequently in terms of $ \vec q_m$ and $\vec \pi_m$) and we see that the both sets of generators (except Hamiltonian) agree at time $t=0$. This result becomes even more evident if we apply the algorithm of constructing integrals of motion for Hamiltonian system with symmetry presented in Ref. \cite{b14}. Namely, for the Lie algebra spanned by $X_i$, $i=1,\ldots, n$, $[X_i,X_j]=\sum_{k=1}^nc_{ij}^kX_k$, with the adjoint action \be \label{e53} Ad_g(X_i)=gX_ig^{-1}=\sum_{j=1}^n D^j_i(g)X_j, \ee the integrals of motion $X_i(\xi,t)$ corresponding to the generators $X_i$ are of the form \be \label{e54} X_i(\xi,t)=\sum_{j=1}^nD_i^j(e^{tH})\xi_j, \ee where $\xi's$ are the coordinates of the dual space to the Lie algebra (more precisely, their restriction to the orbits of the coadjoint action in the dual space). \par Let us apply this approach to our case. One can check that for $H,D,K,J^{\alpha\beta}$ Eq. (\ref{e54}) gives (\ref{e51}) and (\ref{e52}). For $\vec C_p$ we have \be \label{e55} \vec C_p= e^{tH}\tilde{\vec C}_pe^{-tH}=\sum_{r=0}^{2n-1}a_{pr}(t)\tilde{\vec C}_r, \quad p=0,\ldots,2n-1, \ee where the functions $a_{pr}$ satisfy the set of equations \be \label{e56} \dot a_{pr}(t)=(r+1)a_{p,r+1}(t)+(r-2n)\omega^2a_{p,r-1}(t), \ee with $a_{k,-1}=a_{k,2n}=0$ and the initial conditions $a_{pr}(0)=\delta_{pr}$ . Substituting $a_{pr}(t)=\hat a_{pr}(t\omega)\omega^r$ we obtain \be \label{e57} \dot {\hat a}_{pr}(t)=(r+1)\hat a_{p,r+1}(t)+(r-2n)\hat a_{p,r-1}(t), \ee with appropriate initial conditions. It turns out that for fixed $p$ equation (\ref{e57}) is strongly related to the evolution of ${\vec q}$'s and $\vec \pi$'s in the PU model with odd frequencies. More precisely, the canonical equations of motion for the Hamiltonian (\ref{e48}) are equivalent to Eq. (\ref{e57}) for fixed $p$ (cf., \cite{b17}). Consequently, the solution can be written in terms of combinations of harmonics with odd frequencies: \be \label{e58} \hat a_{pr}(t)=\sideset{}{'}\sum_{a=-(2n-1)}^{2n-1}i^r\beta_{ra}e^{iat}s_{a}^p, \ee where $\beta_{pa}$ is given by (\ref{a23}) and $s^p_a$ are some constants (see, \cite{b17})). Taking into account the initial conditions, we obtain \be \label{e59} a_{pr}(t)=\omega^{r-p}\sideset{}{'}\sum_{a=-(2n-1)}^{2n-1}i^r\beta_{ra}\gamma_{ap}e^{iat\omega}. \ee By virtue of Eqs. (\ref{a24}) and (\ref{a28}), we have \be \label{e60} \begin{split} a_{pr}(t)=\omega^{r-p}\sum_{k=1}^{n}\beta_{rk}^\pm\gamma_{kp}^\pm \cos {(2k-1)\wt}, \end{split} \ee where upper (lower) sign corresponds to $p,r$ even (odd); and \be \label{e61} \begin{split} a_{pr}(t)=\mp\omega^{r-p}\sum_{k=1}^{n}\beta_{rk}^\mp\gamma_{kp}^\pm \sin {(2k-1)\wt}, \end{split} \ee where upper (lower) sign corresponds to $p$ even and $r$ odd ($p$ odd and $r$ even). Having the explicit form of $a_{pr}(t)$, and using Eqs. (\ref{e47}) and (\ref{e55}) we obtain $\vec C's$ in terms of $\vec q$ 's and $\vec \pi$'s: \be \label{e62} \vec C_p=\sum_{r=0}^{n-1}\left((-1)^{r-1}r!a_{pr}(t)\vec \pi_r+(2n-1-r)!a_{p,2n-1-r}(t)\vec q_r\right). \ee As we have mentioned above (\ref{e62}) is related by canonical transformation (\ref{e49}) to (\ref{e23}) and (\ref{e23a}). \section{Discussion} Let us summarize. In the present paper we focused on the Hamiltonian approaches to the PU model and its symmetries. First, we derived the form of the symmetry generators, in the original Pais and Uhlenbeck approach (for both generic and odd frequencies). We have shown that the resulting algebra is the central extension of the one obtained on the Lagrangian level, i.e., the centrally extended $l$-conformal Newton-Hooke algebra in the case of odd frequencies and the algebra defined by Eqs. (\ref{e5}) and (\ref{e7}), in the generic case. Next, we considered the Ostrogradski method of constructing Hamiltonian formalism for theories with higher derivatives. We derived the canonical transformation (Eqs. (\ref{e40})-(\ref{e41})) leading the Ostrogradski Hamiltonian to the one in decoupled oscillators approach. \par Let us note that the both approaches, mentioned above, do not distinguish the odd frequencies and in that case do not uncover the richer symmetry. A deeper insight is attained by nothing that for odd frequencies an alternative Hamiltonian formalism can be constructed. It is based on the Hamiltonian formalism for the free higher derivatives theory exhibiting the $l$-conformal Galilei symmetry. More precisely, we add to the Hamiltonian of the free theory the conformal generator. As a result, we obtain the new Hamiltonian, which turns out to be related, by canonical transformation (\ref{e49}), to the PU one. This construction can be better understood from the orbit method point of view, where the construction of dynamical realizations of a given symmetry algebra is related to a choice of one element of the dual space of the algebra as the Hamiltonian (see, \cite{b14} and the references therein). In our case, both algebras ($l$-Galilei and $l$-Newton-Hooke) are isomorphic to each other; only the one generator, corresponding to the Hamiltonian, differ by adding the conformal generator of the free theory. This gives the suitable change in the dual space and consequently the definition (\ref{e48}). \par The change of the Hamiltonian alters the dynamics, which implies different time dependence of the symmetry generators (which do not commute with $H$); however, all PU generators should be expressed in terms of the generators of the free theory (for $t=0$). This fact was confirmed by applying the method presented in Ref. \cite{b14} as well as, directly, by the canonical transformation (\ref{e49}) to the decoupled oscillators approach for the PU model. \par Turning to possible further developments, let us recall that in the classical case ($l=\frac 12$) the dynamics of harmonic oscillator (on the half-period) is related to the dynamics of free particle by well known Niederer's transformation \cite{b18} (this fact has also counterpart on the quantum level). It turns out that this relation can be generalized to an arbitrary half-integer $l$ \cite{b11} on the Lagrangian level; on the Hamiltonian one, we encounter some difficulties since there is no straightforward transition to the Hamiltonian formalism for a theory with higher derivatives. However, in the recent paper \cite{b19} the canonical transformation which relates the Hamiltonian (\ref{e48}) to the one for free theory (the first line of (\ref{e47})) has been constructed; it provides a counterpart of classical Niederer's transformation for the Hamiltonian formalism developed in Section 5. Using our results one can obtain similar transformation for both remaining Hamiltonian approaches. We also believe that the results presented here can help in constructing quantum counterpart of the Niederer's transformation for higher $l$ as well as to study of the symmetry of the quantum version of PU oscillator. \vspace{0.5cm} \par {\bf Acknowledgments.} The author is grateful to Joanna Gonera, Piotr Kosi\'nski and Pawe\l\ Ma\'slanka for useful comments discussions. The e-mail discussion with Professors Anton Galajinsky and Ivan Masterov is highly acknowledged. The work is supported by the grant of National Research Center number DEC-2013/09/B/ST2/02205.
\section{Detector Borexino} Borexino is a unique detector able to perform measurement of solar neutrinos fluxes in the energy region around 1 MeV or below because of its low level of radioactive background. After several years of efforts and tests with the prototype CTF detector the design goals have been reached and for some of the radioactive isotopes (internal $^{238}$U and $^{232}$Th) largely exceeded. The low background is an essential condition to perform the measurement: in fact solar neutrinos induced scintillations cannot be distinguished on an event by event analysis from the ones due to background. The energy shape of the solar neutrino is the main signature that has to be recognized in the experimental energy spectrum by a suitable fit procedure that includes the expected signal and the background. The basic signature for the mono-energetic 0.862 MeV $^{7}$Be neutrinos is the Compton-like edge of the recoil electrons at 665 keV. The detector is located deep underground (approximately 3800 m of water equivalent, mwe) in the Hall C of the Laboratori Nazionali del Gran Sasso (Italy), where the muon flux is suppressed by a factor of 10$^{6}$. The main goal of the experiment was the detection of the monochromatic neutrinos that are emitted in the electron capture of $^{7}$Be in the Sun with 5\% precision. \begin{figure} \begin{centering} \includegraphics[width=0.55\paperwidth,height=0.4\paperwidth]{Figures/Detector.pdf} \par\end{centering} \caption{\label{Detector}Sketch of the Borexino detector. The base of the dome-like structure is 18 m in diameter.} \end{figure} The complete up to date technical description of the Borexino detector has been reported in \cite{B02b} and \cite{Brx08}. The detector is schematically depicted in Fig.\ref{Detector}. The inner part is an unsegmented stainless steel sphere (SSS) that is both the container of the scintillator and the mechanical support of the photomultipliers. Within this sphere, two nylon vessels separate the scintillator volume in three shells of radii 4.25 m, 5.50 m and 6.85 m, the latter being the radius of the SSS itself. The inner nylon vessel (IV) contains the liquid scintillator solution, namely PC (pseudocumene, 1,2,4-trimethylbenzene C$_{6}$H$_{3}$(CH$_{3}$)$_{3}$) as a solvent and the fluor PPO (2,5- diphenyloxazole, C$_{15}$H$_{11}$NO) as a solute at a concentration of 1.5 g/l (0.17\% by weight). The second and the third shell contain PC with a small amount (5 g/l) of DMP (dimethylphthalate) that is added as a light quencher in order to further reduce the scintillation yield of pure PC. The PC/PPO solution adopted as liquid scintillator satisfies specific requirements: high scintillation yield ($\sim10^{4}$ photons/MeV), high light transparency (the mean free path is typically 8 m) and fast decay time ($\sim$3 ns), all essential for good energy resolution, precise spatial reconstruction, and good discrimination between $\beta$-like events and events due $\alpha$ particles. Furthermore, several conventional petrochemical techniques are feasible to purify the hundred of tons of fluids needed by Borexino. The scintillation light is collected by 2212 photomutipliers (PMTs) that are uniformly attached to the inner surface of the SSS. All but 384 photomultipliers are equipped with light concentrators that are designed to reject photons not coming from the active scintillator volume, thus reducing the background due to radioactive decays originating in the buffer liquid or $\gamma$'s from the PMTs. The tank has a cylindrical base with a diameter of 18 m and a hemispherical top with a maximum height of 16.9 m. The Water Tank (WT) is a powerful shielding against external background ($\gamma$--rays and neutrons from the rock) and is also used as a Cherenkov muon counter and muon tracker. The muon flux, although reduced by a factor of $10^{6}$ by the 3800 m.w.e. depth of the Gran Sasso Laboratory, is of the order of 1 m$^{-2}$ h$^{-1}$, corresponding to about 4000 muons per day crossing the detector. This flux is well above Borexino requirements and a strong additional reduction factor (about $10^{4}$) is necessary. Therefore the WT is equipped with 208 photomultipliers that collect the Cherenkov light emitted by muons in water. In order to maximize the light collection efficiency the SSS and the interior of the WT surface are covered with a layer of Tyvek, a white paper-like material made of polyethylene fibers. The Borexino has an excellent energy resolution for its size, this is the result of the high light yield of $\sim500$ p.e./MeV/2000 PMTs. The energy resolution (1$\sigma$) at the $^{7}$Be Compton edge energy (662 keV) is as low as 44 keV (or 6.6\%). \section{Solar neutrino measurements at the first stage of the experiment} Analysis of the Borexino data showed that the main goals concerning the natural radioactivity have been achieved. The contamination of the liquid scintillator with respect to the U/Th is at the level of ${10}^{-17}$ g/g; the contamination with $^{40}$K is at the level of $10^{-19}\;$g/g ($10^{-15}$ in natural potassium); the $^{14}$C content is $2.7\pm0.1\times10^{-18}$ g/g with respect to the $^{12}$C. Among the other contamination sources only $^{85}$Kr, $^{210}$Bi and $^{210}$Po have been identified. The $^{85}$Kr counts $\sim$0.3 ev/day/tone, it is $\beta$- emitter with 687 keV end-point. The $^{210}$Po is the most intense contamination (with initial activity of 60 counts/day/tone), it decays emitting monoenergetic $\alpha$ with 5.3 MeV energy, the half-life time of the isotope is 134 days. The residual contaminations do not obscure the expected neutrino signal, the presence of the 862 keV monoenergetic $^{7}$Be solar neutrino is clearly seen in the experimental spectrum. In such a way, the collaboration succeeded to purify the liquid scintillator from residual natural radioactive isotopes down to the levels much lower than was initially envisaged for the $^{7}$Be neutrino measurement, which resulted in broadening of the initial scientific scope of the experiment. The main goal of the experiment was the detection of the monochromatic neutrinos that are emitted in the electron capture of $^{7}$Be in the Sun with 5\% precision. This goal has been achieved during the first stage of the experiment \cite{BOR07,Be708,Be7}. The Borexino reported the first measurement of neutrino $^{8}$B neutrinos with liquid scintillator detector with a 3 MeV threshold on electrons recoil \cite{B8}. The stability of the detector allowed also to study the day-night effect of the $^{7}$Be solar neutrino signal, thereby allowing to completely exclude the LOW solution of the neutrino oscillation based on solar data alone \cite{DN11}. Finally, the low background of the detector, the refined analysis on threefold coincidences \cite{CNO06} and the positronium discrimination method based on the positronium formation study made it possible to explore the 1-2 MeV region with unprecedented sensitivity. This led to the first observation of solar neutrinos from the basic pep reaction \cite{PEP11}. In addition, the best limit for the CNO production in a star has been established. In this way, Borexino has completed direct detection of Be-7, pep and B-8 solar neutrino components thereby providing complete evidence of the transition from MSW and vacuum oscillation of the LMA solution of the Solar Neutrino Problem (fig. \ref{Pee}). \begin{figure} \begin{centering} \includegraphics[width=0.6\textwidth,height=0.5\textwidth]{Figures/Pee} \par\end{centering} \caption{\label{Pee}electron neutrino survival probability, three points in this plot are derived from the Borexino data. It should be noted that the precision of the Be-7 neutrino flux itself is much better than the Pee error at the corresponding point, this is related to the solar model uncertainties. } \end{figure} \section{Solar neutrino program of the Borexino Phase II} One of the goal of the Borexino experiment is the measurement of all the solar neutrino fluxes, with the exception of the hep flux, too faint for detection in Borexino. The $^{7}$Be, pep and $^{8}$B (this last with the lowest threshold to date) have been already measured, but the experimental uncertainties can be reduced. In addition Borexino will try to measure the pp and CNO fluxes. \begin{table} \begin{tabular}{|c|c|c|c|c|} \hline $\nu$ flux & GS98 & AGS09 & cm$^{-2}$s$^{-1}$ & Experimental result\tabularnewline \hline \hline pep & $1.44\pm0.012$ & $1.47\pm0.012$ & $\times10^{8}$ & 1.6$\pm0.3$ Borexino\tabularnewline \hline $^{7}$Be & $5.00\pm0.07$ & $4.56\pm0.07$ & $\times10^{9}$ & $4.87\pm0.24$ Borexino\tabularnewline \hline $^{8}$B & $5.58\pm0.14$ & $4.59\pm0.14$ & $\times10^{6}$ & $5.2\pm0.3$ SNO+SK+Borexino+KamLAND\tabularnewline & & & & $5.25\pm0.16_{-0.013}^{+0.011}$SNO-LETA\tabularnewline \hline $^{13}$N & $2.96\pm0.14$ & $2.17\pm0.14$ & $\times10^{8}$ & \tabularnewline \cline{1-4} $^{15}$O & $2.23\pm0.15$ & $1.56\pm0.15$ & $\times10^{8}$ & $<$7.4 Borexino (total CNO)\tabularnewline \cline{1-4} $^{17}$F & $5.52\pm0.17$ & $3.40\pm0.16$ & $\times10^{8}$ & \tabularnewline \hline \end{tabular} \caption{\label{SSMvsData}SSM predictions and current experimental results} \end{table} In table \ref{SSMvsData} the solar fluxes measured by Borexino so far are compared with the SSM prediction, for low and high metallicity. The experimental results agree, within the errors, with the SSM predictions, but cannot distinguish between the two metallicities, due to the uncertainties of the model and the experimental errors. It would be useful, at this moment, to recall what the metallicity puzzle is. The solar surface heavy element abundance has been calculated about ten years ago with a 1D model, which uses data from spectroscopic observations of the elements present in the photosphere (GS98 \cite{Grevesse}). This model agrees with the helioseismology observations, namely the measurement of the speed of the mechanical waves in the Sun. More recently a 3D hydro-dynamical model (AGSS09 \cite{Asplund}) of the near-surface solar convection, with improved energy transfer, has changed the Z/X ratio with respect to the previous 1D treatment: 0.0178 (low metallicity) to be compared with the previous 0.0229 (high metallicity). The 3D model results perfectly reproduce the observed solar atmospheric line (atomic and molecular) profiles and asymmetries, but are in clear disagreement with the helioseismology data. At present there is no satisfactory solution to this controversy \cite{Serenelli}. The 1D and the 3D models predict different neutrino fluxes from the various nuclear reactions, as shown in Table 1, where they are compared with the experimental results obtained until now. As stated above, it is not possible, at present, to discriminate the solutions due to model uncertainties and experimental errors. A measurement of the CNO flux, with reasonable errors, could distinguish between the two models which predict substantially different fluxes. The pp solar neutrino flux has been never measured directly. Gallex and Sage have measured the integrated solar flux from 233 keV, which, together with the Borexino $^{7}$Be neutrino flux measurement and the experimental data on the $^{8}$B neutrino flux, can be used to infer the pp neutrino flux with a relatively small uncertainty, once the luminosity constraint is applied (fig. \ref{Pee}). Nevertheless a direct experimental observation, which can be compared with the solar luminosity and the SSM prediction, would be an important achievement. The pp flux measurement is part of the Borexino phase 2 program. \subsection*{Improvement of the$^{7}$Be solar neutrino flux measurement} Improving the $^{7}$Be flux measurement is one of the goals for Borexino phase II. The physics goals of this study can be summarized in three main points. \begin{enumerate} \item Reduction of the total error (statistical+systematic) down to hopefully $\sim$3\%. Even with such a precision this measurement cannot solve unambiguously the metallicity puzzle, because of the uncertainties of the SSM. The $^{7}$Be flux measured in phase I falls in between the two predicted flux values, for high and low metallicities, and a smaller uncertainty would not help if this will be the case also for phase II. A very precise experimental determination of the $^{7}$Be flux remains nonetheless an important tool for testing the Solar Model as well as a remarkable technical achievement. \item In the context of the neutrino physics, the Non Standard neutrino Interactions are currently debated. One way to study them is to analyze the shape of the oscillation vacuum-matter transition region. While $^{7}$Be cannot have a conclusive role in this matter, it can nevertheless help in restricting the range of the NSI flavor diagonal terms. \item It is possible to constraint the NSI parameters studying the $\nu-e$ elastic scattering. Bounds are imposed by various other experiments on solar, atmospheric and reactor (anti)neutrinos. But $^{7}$Be neutrinos have the strong advantage of being mono-energetic ($^{8}$B neutrino detected by the other solar experiments in real time have a continuous energy spectrum). In Borexino, the limitation to this analysis comes from the residual background, especially $^{85}$Kr, and, to lesser extend, $^{210}$Bi, which can mimic non-zero values of the NSI parameters $\epsilon_{\alpha L}$ and $\epsilon_{\alpha R}$. An increase of statistics does not help much if not accompanied by a reduction of such background. \end{enumerate} \subsection*{pp Solar Neutrino measurement} This is the most important target of opportunity for Phase 2. The very low $^{85}$Kr and reasonably low $^{210}$Bi achieved, make a direct pp measurement a reality. A careful understanding of the spectrum response in the $^{14}$C end-point region is crucial, its study is possible through a dedicated effort. The main problem in the pp-neutrino study is the disentanglement of the very tail of the $^{14}$C spectrum (with possible pile-up) from the pp- neutrino spectrum. The feasibility of the measurement is under study. A direct detection of pp neutrinos would be a spectacular result and would justify the phase II alone. The analysis group performed a study of sensitivity to pp solar neutrinos with the current background levels achieved, the expected statistical precision of the pp-neutrino flux measurement is below 10\%, the systematics error will be mainly connected with uncertainty of the $^{14}C$ pile-up spectrum and intensity determination. \subsection*{pep Solar Neutrino measurement} The first indication for pep solar neutrinos has been reported by the collaboration \cite{PEP11}. The value for the pep interaction rate obtained in Phase I (590 live-days) was $3.1\pm0.6(stat)\pm0.3(syst)$ cpd/day/100 tones, the absence of a pep signal was rejected at 98\% C.L. The current measurement, in conjunction with the SSM (the uncertainty in the pep flux is as low as 1.2\%), yields a survival probability of Pee = $0.62\pm0.17$ though the uncertainties are far from Gaussian. The precision is dominated by the statistical uncertainty (about 20\%), though with more (background-free) data, systematic uncertainties (10\%) will start to become important. This is an extremely important result, but shy of the first measurement of pep solar neutrinos. The addition of a modest batch of data with $^{210}$Bi reduced at or below 30 counts/(100 ton $\times$ day) will result in the first measurement (3$\sigma$) of pep solar neutrinos. A much prolonged data taking could also result in a 5$\sigma$ precision measurement. The measurement will allow to gauge the survival probability in the immediate proximity of the transition between two different oscillation regimes. \subsection*{solar $^{8}$B neutrino flux measurements.} The Borexino detector is the first large volume liquid scintillator detector sensitive to the low-energy solar neutrinos. It possesses a very good energy resolution in comparison to the water Cherenkov detectors, what allows to search for the solar $^{8}$B neutrinos starting practically from the energies of the so called Thallium limit (maximum energy of $\gamma$ rays from the chains of radioactive decay of $^{232}$Th and $^{238}$U; gamma-quantum with maximum energy E=2.6 MeV is emitted in the decay of $^{208}$Tl). The measurements of the $^{8}$B above 2.8 MeV has been performed using one year statistics (246 days of live- time) of the Borexino data \cite{B8}. The threshold of 2.8 MeV is the lowest achieved so far in the $^{8}$B neutrino real-time measurements. The interest in the neutrino flux measurement with low threshold comes from the peculiar properties of the survival probability in this energy region. The electron neutrino oscillations at E<2 MeV are expected to be driven by the so called vacuum oscillation, and at energies E>5 MeV - by resonant matter-enhanced mechanism. The energy region in between has never been investigated in spectrometric regime, and is of particular interest because of the expected smooth transition between the two types of oscillations. The rate of $^{8}$B solar neutrino interaction as measured through their scattering on the target electrons is 0.22$\pm$0.04(stat)$\pm$0.01(sys) cpd/100 tons. This corresponds to an equivalent electron neutrino flux of $\Phi_{^{8}\text{B}}^{ES}$=(2.4$\pm$0.4$\pm$0.1)$\times$10$^{6}$ cm$^{-2}$s$^{-1}$, as derived from the elastic scattering only, in good agreement with existing measurements and predictions. The corresponding mean electron neutrino survival probability, assuming the BS07(GS98) Standard Solar Model (High Z model), is 0.29$\pm$0.10 at the effective energy of 8.6 MeV. The ratio between the measured survival probabilities for $^{7}$Be and $^{8}$B is 1.9$\sigma$ apart from unity (see Fig.\ref{Pee}). For the first time the presence of a transition between the low energy vacuum-driven and the high-energy matter-enhanced solar neutrino oscillations is confirmed using the data from a single detector, the result is in agreement with the prediction of the MSW-LMA solution for solar neutrinos. Acquiring more statistics (of up to 5 years of the calendar time) the Borexino will provide the competitive measurement of the $^{8}$B neutrino flux. \subsection{SOX campaign} The Borexino large size and possibility to reconstruct an interaction point (with a precision of 14~cm at 1 MeV energy deposit) makes it an appropriate tools for searching of sterile neutrinos. If the oscillation baseline is about 1 m (which corresponds to $\Delta m^{2}\sim$1 eV$^{2}$), exposure of the detector to a compact powerful neutrino source should give rise to a typical oscillation picture with dips and rises in the spatial distribution of events density with respect to the source. Right beneath the Borexino detector, there is a cubical pit (side 105 cm) accessible through a small squared tunnel (side 95 cm) that was built at the time of construction with the purpose of housing neutrino sources for calibration purposes. Using this tunnel, the experiment with neutrino source can be done with no changes to the Borexino layout. The center of the pit is at 8.25 m from the detector center, requiring a relatively high activity of the neutrino source in order to provide detectable effect. The experiment SOX (for Short distance neutrino Oscillations with BoreXino) \cite{SOX} will be carried in three stages with gradually increasing sensitivity: \begin{description} \item[Phase A] a $^{51}$Cr neutrino source of 200-400 PBq activity deployed at 8.25 m from the detector center (external with respect to the detector); \item[Phase B] deploying a $^{144}$Ce-$^{144}$Pr antineutrino source with 2-4 PBq activity at 7.15 m from the detector center (placed in the detector's water buffer); \item[Phase C] a similar $^{144}$Ce-$^{144}$Pr source placed right in the center of the detector. \end{description} \begin{figure}[t] \begin{center} \includegraphics[width=0.80\textwidth]{Figures/DetectorSources.pdf} \caption{\label{fig:detector} Layout of the Borexino detector and the approximate location of the neutrino and anti-neutrino sources in the three phases.} \end{center} \end{figure} Figure 1 shows a schematic layout of the Borexino detector and the approximate location of the neutrino and anti-neutrino sources in the three phases. Two types of neutrino sources are considered: the $^{51}$Cr source and the $^{144}$Pr based source. $^{51}$Cr decays via electron capture into $^{51}$V, emitting two neutrino lines of 750 keV (90\%) and 430 keV (10\%), while $^{144}$Pr decays $\beta$ into $^{144}$Nd with an end--point of 3 MeV (parent $^{144}$Ce decays too, but end-point of its $\beta$-decay is below the IBD threshold). The portion of the $^{144}$Pr spectrum above the 1.8 MeV detection threshold is the only of importance for the experiment. Elastic scattering of $\bar{\nu}_e$ on electrons induce negligible background. The source activity of 200-400 PBq is challenging, but only a factor 2-4 higher than what already done by Gallex and SAGE in the 90's. The $^{144}$Ce--$^{144}$Pr experiment in Phases B and C doesn't require high source activity. The Phase C is the most sensitive but it can be done only after the shutdown of the solar neutrino program, because it needs modification of the detector. The Phases A and B will not disturb the solar neutrino program of the experiment, which is supposed to continue until the end of 2015, and do not require any change to Borexino hardware. The challenge for the Phase C is constituted by the large background induced by the source in direct contact with the scintillator, that can be in principle tackled thanks to the correlated nature of the $\bar{\nu}_e$ signal detection. In Phase B this background, though still present, is mitigated by the shielding of the buffer liquid. Borexino can study short distance neutrino oscillations in two ways: by comparing the detected number of events with expected value (disappearance technique, or total counts method), or by observing the oscillation pattern in the events density over the detector volume (waves method). In the last case the typical oscillations length is of the order The variations in the survival probability $P_{ee}$ could be seen on the spatial distribution of the detected events as the waves superimposed on the uniformly distributed background. Oscillation parameters can be directly extracted from the analysis of the waves. The result may be obtained only if the size of the source is small compared to the oscillation length. The $^{51}$Cr source will be made by about 10-35 kg of highly enriched Cr metal chips which have a total volume of about 4-10 l. The source linear size will be about 15-23 cm, comparable to the spatial resolution of the detector. The $^{144}$Ce--$^{144}$Pr source is even more compact. All simulations shown below takes into account the source size. In Phase A the total counts method sensitivity is enhanced by exploiting the fact that the life-time of the $^{51}$Cr is relatively short. In Phases B and C this time-dependent method is not effective because the source life-time is longer (411 days), but this is compensated by the very low background and by the larger cross-section. The total counts and waves methods combined together yield a very good sensitivity for both experiments. Besides, the wave method is independent on the intensity of the source, on detector efficiency, and is potentially a nice probe for un-expected new physics in the short distance behavior of neutrinos or anti-neutrinos. The sensitivity of SOX with respect to oscillation into sterile neutrino was evaluated with a toy Monte Carlo. Expected statistical samples (2000 events) were generated for each pair of oscillation parameters. We assume a period of 15 weeks of stable data taking before the source insertion in order to accurately constrain the background. The background model includes all known components, identified and accurately measured during the first phase of Borexino. We built the confidence intervals from the mean $\chi^2$ for each couple of parameters with respect to the non-oscillation scenario. The result is shown in Fig.~\ref{fig:sensitivity}, as one can see fro the figure the reactor anomaly region of interest is mostly covered. \begin{figure}[t] \begin{center} \includegraphics[width=0.95\textwidth]{Figures/SOX_allphases.pdf} \end{center} \caption{\label{fig:sensitivity} Sensitivity of the Phase A ($^{51}$Cr external, blue), of Phase B ($^{144}$Ce--$^{144}$Pr external, red) and Phase C ($^{144}$Ce--$^{144}$Pr center, green). The gray area is the one indicated by the reactor anomaly, if interpreted as oscillations to sterile neutrinos. Both 95\% and 99\% C.L. are shown for all cases. The yellow line indicates the region already excluded in \cite{bib:palazzo2}.} \end{figure} The simulation for the Phase A are shown for a single irradiation of the $^{51}$Cr source up to the initial intensity of 370 PBq (10 MCi) at the site. A similar result can be obtained with two irradiations of about 200 PBq if higher intensity turns out to be beyond the technical possibilities. The single irradiation option is preferable and yields a slightly better signal to noise ratio. The physics reach for the $^{144}$Ce--$^{144}$Pr external (Phase B) and internal (Phase C) experiments, assuming 2.3 PBq (75 kCi) source strength and one and a half year of data taking) is shown in the same figure (Fig.~\ref{fig:sensitivity}). The $\chi^2$ based sensitivity plots are computed assuming significantly bigger volume of liquid scintillator (spherical vessel of 5.5 m radius), compared to the actual volume of liquid scintillator (limited by a sphere with 4.25 m radius) used for the solar phase. Such an increase will be made possible by the addition of the scintillating fluor (PPO) in the inner buffer region (presently inert) of the detector. We have also conservatively considered exclusion of the innermost sphere of 1.5 m radius from the analysis in order to reject the gamma and bremsstrahlung backgrounds from the source assembly. Under all these realistic assumptions, it can be noted from Fig.~\ref{fig:sensitivity} that the intrinsic $^{144}$Ce--$^{144}$Pr sensitivity is very good: for example the 95\% C.L. exclusion plot predicted for the external test covers adequately the corresponding reactor anomaly zone, thus ensuring a very conclusive experimental result even without deploying the source in the central core of the detector. \section{Conclusions} Borexino achieved its main goal, but is still a challenging detector. Due to achieved level of purification (much higher than it was needed for $^{7}$Be neutrino detection) some measurements of solar neutrino fluxes beyond the original program were performed. Borexino-II is operating with repurified LS, at new levels of radiopurity, two years of data are collected and are being analyzed aiming the pp-neutrino flux and CNO neutrino flux measurement. Borexino is also an ideal detector to test the sterile neutrinos through the disappearance/wave effects. The proposed staged approach ($^{51}$Cr source at the first stage and two $^{144}$Ce–$^{144}$Pr experiments) is a comprehensive sterile neutrino search which will either confirm the effect or reject it in a clear and unambiguous way. In particular, in case of one sterile neutrino with parameters corresponding to the central value of the reactor anomaly, SOX will surely discover the effect, prove the existence of oscillations and measure the parameters through the “oscillometry” analysis. \section*{Acknowledgments} Borexino was made possible by funding from INFN (Italy), NSF (USA), BMBF, DFG, and MPG (Germany), NRC Kurchatov Institute (Russia), MNiSW (Poland, Polish National Science Center (grant DEC-2012/06/M/ST2/00426)), Russian Foundation for Basic Research (Grant 13-02-92440 ASPERA, the NSFC-RFBR joint research program) and RSCF research program (Russia). We acknowledge the generous support of the Gran Sasso National Laboratories (LNGS). SOX is funded by the European Research Council.
\section{Introduction} Cells often migrate in response to external signals, including chemical and mechanical signals. Thereby the interfacial growth of filamentous actin polymer networks plays an important role \cite{SvBo99,LBP99}. For example, cell crawling on a two-dimensional substrate involves the formation of a cytoplasmic membrane protrusion pointing in the direction of motion. Thereby, the necessary force for extending the membrane is provided by the polymerization of actin, a process far from chemical equilibrium, which converts chemical into mechanical energy. The same molecular machinery is also responsible for the propulsion of cellular organelles \cite{TRM00}, pathogens \cite{YTA99,CCG95} or biomimetic objects, such as spherical beads \cite{OuT99,NGF00,Bernheim_Science_2002,GPP05,DSR08,AMMG10}, vesicles \cite{UCA03,GFT03,DSR08}, droplets \cite{BCJ04}, and ellipsoids \cite{LSZT12}. In contrast, cell motion in a three-dimensional substrate is rather driven by blebbing, which relies on the contraction of the actin cortex by myosin motors to form protrusions \cite{Charras_NatRevMolCellBiol_2008}. On the time scale of actin filament growth ($\sim 1\,$s) actin networks behave as nonlinear elastic solids \cite{GSMM04,JMRL07}. Typically, the linear elastic modulus of actin networks formed during the propulsion of pathogens or biomimetic objects is in the range of 1 to 10\,kPa \cite{GCR00,PRFH12}. This raises the question of the coupling mechanism between growth dynamics and deformations (or stresses) in the network, a property which is often either neglected \cite{DALR09,OSS08,Kawska_PNAS_2012} or included via ad hoc assumptions, which do not necessarily respect all symmetries in the system \cite{SPJ04,GPP05}. Our objective is therefore to derive macroscopic evolution equations of actin networks combining a macroscopic constitutive law for its mechanical behavior and the actin polymerization kinetics in a rigorous way. The general formulation we present can be adapted to diverse situations relevant for actin dynamics. As a proof of principle we treat here the case of an elastic actin network growing from a spherical surface, where a spontaneous symmetry-breaking and the onset of motion has been observed experimentally \cite{OuT99,NGF00,Bernheim_Science_2002,GPP05,DSR08,AMMG10}. To briefly outline the experimental observations (more details can be found in \cite{Blanchoin_PhysRev_2014}), actin polymerizes on the surface of a sphere with a radius of about 1\,$\mu$m into a cross-linked/\-entangled network, which forms initially a closed spherical shell. Growth and cross-linking is restricted to a zone close to the surface of the sphere. Monomers diffuse freely through the network to reach the growth zone and are inserted between the already existing network and the sphere, which leads to a relative motion between older network layers and the surface and to the buildup of mechanical stresses. Eventually the actin shell undergoes a spontaneous symmetry-breaking, leading to the formation of an actin tail. Two mechanisms for symmetry-breaking have been proposed: (i) the external actin shell ruptures at one point due to elongational stresses \cite{GPP05}, or (ii) the instability is due to actin polymerizing slower on one side of the surface than on the other side \cite{DSR08}. Our understanding of these processes has been advanced through several theoretical studies based on discrete models on the scale of the filament \cite{OuT99,MoO03}, mesoscopic models \cite{DALR09,OSS08,Kawska_PNAS_2012} or phenomenological continuous models \cite{SPJ04,LLK05,GPP05,JPK08}. Rather surprisingly, a general growth law based on first thermodynamic principles which links the stresses in the network (which depend on the growth history itself and the relevant boundary conditions) to the interface dynamics is lacking. Our approach closes this gap. In this brief exposition, we exploit the framework to show that an instability arises from the interplay between interfacial growth kinetics and mechanical stresses, which reflect the growth history of the network. Thereby the nature of the contact formed between the network and the bead surface (fixed vs. sliding) is crucial in the overall macroscopic dynamics. In a spherical geometry a spontaneous polarization into front and back is possible when the filaments can slide on the surface, but not when they are fixed to the surface. Furthermore we derive scaling laws for the instability characteristics which form a consistent picture with experiments. The problem of symmetry-breaking has been studied theoretically on the continuum level in \cite{SPJ04,JPK08}. However, in contrast to \cite{SPJ04,JPK08} we describe here the buildup of stresses and strains in the network due to growth and the subsequent mechano-chemical coupling in a rigorous way, consistent with a hyperelastic macroscopic constitutive law of the network. For example, the model in \cite{SPJ04} tacitly assumes a vanishing Poisson ratio and the stress distribution can only be solved consistently in an axisymmetric or spherical configuration. The model in \cite{JPK08} is limited to small deformations arising from a small displacment at the internal bead/network interface. Moreover, our description distinguishes between reference and deformed network configurations, which is essential for describing a growing interface in contact with a solid stationary substrate (the spherical surface), where the growth process manifests itself as a displacement of a free interface (in contact with the solvent). This crucial distinction between reference and deformed frame is absent in \cite{SPJ04,JPK08} and limits their predictive power. \section{Model}{ \subsection{Mechanical description of the network} Recently we have proposed a macroscopic mechanical model of actin networks starting from a microscopic description and have shown that it captures the basic bulk rheological properties of actin networks \cite{John_PhysRevE2013}. A major issue is now, how to properly combine interfacial growth and mechanics. We first briefly recall the description of the network mechanics and then introduce the dynamical equations for the interfaces. For simplicity, we consider a 2D geometry (albeit a 3D study does not pose a specific challenge, but increases the technical complexity). We assume a structurally periodic planar filament network with the topology shown in Fig. \ref{network:topo}, which is in contact with the surface of a cylinder of radius $R_0$. The typical scale of the elementary cell $l_0$ is small compared to a macroscopic scale $R_0$, which introduces the small parameter $\eta=l_0/R_0\ll 1$. \begin{figure} \includegraphics[width=0.7\hsize]{figure1.eps} \caption{Sketch of a small part of the filament network showing the different types of elementary vectors $\mathbf{B}_i$ of one elementary cell labeled $(n_1,n_2)$ in different colors.\mylab{network:topo}} \end{figure} Each elementary cell contains one node, identified by a doublet of integers $(n_1,n_2)$, and three filaments, described by the vectors $\mathbf{B}_i$ with $i=1,2,3$. It has a quadrilateral shape with the filaments $\mathbf{B}_1$ and $\mathbf{B}_3$ forming the sides and the filament $\mathbf{B}_2$ forming a diagonal. A node is formed by the intersection of six filaments, three of which belong to the same elementary cell as the node, the other three belonging to neighboring elementary cells. For example, the node ($n_1,n_2$) shown in Fig. \ref{network:topo} connects the filaments $\mathbf{B}_1$, $\mathbf{B}_2$, and $\mathbf{B}_3$ of the same node with the filaments $\mathbf{B}_1$ from node ($n_1-1,n_2$), $\mathbf{B}_2$ from node ($n_1,n_2-1$) and $\mathbf{B}_3$ from node ($n_1+1,n_2-1$). This network topology is one of the most simple topologies. Other structures involving several nodes per elementary cell are possible. However, the resulting mechanical equilibrium equations are much more involved, than for simple structures with only one node per elementary cell. In a simple picture the filaments behave as entropic springs \cite{GSMM04} and the elastic energy of the filament $\mathbf{B}_i$ is given by a harmonic potential \begin{equation} f_i={k\over 2 l_0}(l_i-l_0)^2 \mylab{harm:pot} \end{equation} with the equilibrium strand length $l_0$ and the actual strand length $l_i = |\mathbf{B}_i|=\sqrt{\mathbf{B}_i\cdot\mathbf{B}_i}$. We find it a bit more convenient to rewrite ($\ref{harm:pot}$) in a different way which is valid as long as the actual length $l_i$ is close to $l_0$. Indeed, in this case, since $(l_i^2-l_0^2)^2\simeq 4l_0^2 (l_i-l_0)^2$, we can write \begin{equation} f={k\over 8 l_0^3}\sum_i^{3}\left(l_i^2-l_0^2\right)^2. \mylab{en} \end{equation} Once the microscopic model is introduced, we can now write the macroscopic equation. Owing to the fact that $\eta=l_0/R_0\ll 1$ one introduces a set of continuous variables $(\lambda_1,\lambda_2)$ \cite{John_PhysRevE2013}, such that the node positions are approximated by a continuous vector function $\boldsymbol{\phi}(\lambda_1,\lambda_2)$ and the filament vectors $\mathbf{B}_i$ are given by Taylor expansions of $\boldsymbol{\phi}$ up to $\mathcal{O}(\eta)$ \begin{equation} \mathbf{B}_{1,3}=\pm\mathbf{h}_1+{1\over 2}\mathbf{h}_2 \quad\mbox{and}\quad \mathbf{ B}_2=\mathbf{h}_2, \quad\mbox{with}\quad \mathbf{h}_i=\eta\partial_{\lambda_i}\boldsymbol{\phi}. \mylab{bhdef} \end{equation} The surface in contact with the cylinder is called "inner interface" as opposed to the ``outer interface'' in contact with the solvent. In the axisymmetric situation $\lambda_1\in(0,\Lambda_1)$ and $\lambda_2\in(0,2\pi)$ correspond to the radial and angular directions, respectively. The unit vector in the radial direction is given by $\mathbf{\hat r}=\cos{(\lambda_2)} \mathbf{\hat x}+\sin{(\lambda_2)} \mathbf{\hat y}$. Consequenlty, the vectors $\mathbf{h}_1$ and $\mathbf{h}_2$ point into the radial and angular directions, respectively. In the continuum limit the discrete sum of (\ref{en}) over all nodes can be converted into an integral over the continuous variables in the domain $\Omega$ occupied by network, so that the total network energy reads \begin{equation} F[\boldsymbol{\phi}]={1\over \eta^2}\int_\Omega f(\boldsymbol{\phi}) \,d\lambda_1\,d\lambda_2,\mylab{int:en} \end{equation} where by using (\ref{en}) and (\ref{bhdef}), $f$ can be expressed only in terms of the continuous function $\boldsymbol{\phi}$. Since growth occurs on a slow time scale, as compared to the propagation of sound in the actin network, the system can be viewed at each instant at mechanical equilibrium, which corresponds to a minimum of $F$ with respect to a variation of $\boldsymbol{\phi}$. This yields $ 0=\partial_{\lambda_1}\mathbf{T}_1+\partial_{\lambda_2}\mathbf{T}_2 $ (divergence-free stress, as in classical linear elasticity). The stress is defined as $\mathbf{T}_i=\partial f/\partial \mathbf{h}_i$ (the two components of the vectors $\mathbf{T}_i$ define the stress tensor). The $\mathbf{T}_i$ are explicitely given by \begin{eqnarray} \mathbf{T}_1 & = & {k\over l_0^3}\left[ (l_1^2-l_0^2) \mathbf{B}_1- (l_3^2-l_0^2) \mathbf{B}_3\right] \mylab{T1}\\ \mathbf{T}_2 & = & {k\over l_0^3} \left[{(l_1^2-l_0^2)\over 2} \mathbf{B}_1+ (l_2^2-l_0^2) \mathbf{B}_2+{(l_3^2-l_0^2)\over 2} \mathbf{B}_3\right].\mylab{T2} \end{eqnarray} and define the constitutive law (relation between stress and strain). Physically, the vector $\mathbf{T}_1$ ($\mathbf{T}_2$) denotes the force exerted on a facet oriented in the direction $\mathbf{h}_2$ ($\mathbf{h}_1$) with length $|\mathbf{h}_2|$ ($|\mathbf{h}_1|$). The associated boundary conditions are \begin{eqnarray} |\boldsymbol{\phi}| & = & R_0 \quad\mbox{and}\quad \mathbf{T}\cdot\mathbf{t} = 0 \quad\mbox{at the inner inteface \mylab{bc1}\\ \mathbf{T} & = & 0 \quad\mbox{at the outer interface} \mylab{bc2} \end{eqnarray} where $\mathbf{T}$ and $\mathbf{t}$ denote the traction force and the tangent vector at the interface. (\ref{bc1}) is equivalent to a shear free inner interface (where actin filament nucleators are present) which is everywhere in contact with the cylinder surface and (\ref{bc2}) is equivalent to a force free outer interface. It has been shown \cite{John_PhysRevE2013} that the potential energy per node $f$ [Eq.\,(\ref{en})] is equivalent to the strain energy density of an isotropic St.-Venant-Kirchhoff hyperelastic solid with the Lam\'e coefficients $\lambda=\mu=\sqrt{3} k/( 4 l_0)$, the Young's modulus $Y=2k/(\sqrt{3}l_0)$ and the Poisson ratio $\sigma=1/3$. This constitutive law results in a strain stiffening and negative normal forces under a simple volume conserving shear, a typical behavior observed experimentally for semiflexible networks \cite{GSMM04,JMRL07}. \subsection{Growth kinetics and interface dynamics} The main issue that remains to be addressed, is how one could link mechanics with actin growth dynamics in a consistent way, and what are the far-reaching consequences. For simplicity, we consider the case that the network grows at the inner interface. That is, the topology of the network remains unchanged while adding or subtracting an elementary material element at the interface. The cost in energy by adding (subtracting) a material element will define the chemical potential difference $\Delta\mu$ that drives the interface using the following kinetics relating the normal velocity $v_n$ to the chemical potential balance \begin{equation} v_n=-M\Delta\mu=-M(\Delta\mu_c+\Delta\mu_m)\,. \mylab{vn} \end{equation} $M$ is a positive mobility constant. $\Delta\mu$ is composed of an attachment part $\Delta\mu_c$ due to chemical bond formation (we assume it to be a constant) and an elastic part $\Delta\mu_m={\partial F\over\partial \Omega}$, which is given by the functional derivative of the total strain energy $F$ (\ref{int:en}) with respect to a change in the shape of the network by respecting the boundary conditions (\ref{bc1},\ref{bc2}). We merely focus here on the main outcomes (details are in Appendix \ref{der:pot}). For the inner interface we find \begin{equation} \Delta\mu_{mi}= {1\over \eta^2} \left[f-\mathbf{T}\cdot \mathbf{h}\right], \mylab{mumi} \end{equation} where $\mathbf{T}=\mathbf{T}_i\nu_i$ denotes the traction force at the interface with $\nu_i$ being the $ith$-component of the unit normal outward vector in the material frame $(\lambda_1,\lambda_2)$. $\mathbf{h}=\mathbf{h}_i\nu_i$ is a measure for the displacement of the old network interface due to the insertion of new material between the solid cylinder and the soft network. All quantities are calculated at the interface position where the material is added. Intuitively one can interpret Eq. (\ref{mumi}) in the following way. The energy cost for inserting a material element at the internal boundary contains the straining of the material element to the same state as neighboring interfacial elements and also the work which is necessary to displace the old interface against the traction force to make room for the new material (the cylinder being rigid cannot be displaced). Expression (\ref{mumi}) is the general form of the chemical potential difference at the internal interface, which is valid for any material whose constitutive law can be expressed in the form of Eqs. (\ref{T1}) and (\ref{T2}). At the force free external interface the elastic chemical potential is simply given by $\mu_{me}={1 \over \eta^2} f$. Setting $\Delta\mu_m=\Delta\mu_{mi}$ and reporting (\ref{T1}) and (\ref{T2}) into (\ref{mumi}) leads to a nonlinear evolution equation (\ref{vn}) relating growth speed and direction $v_n$ [Eq. (\ref{vn})] to the configuration of the network $\boldsymbol{\phi}$. The resulting equation (\ref{vn}), together with (\ref{T1},\ref{T2}) and boundary conditions (\ref{bc1},\ref{bc2}) constitute the general framework that can be applied to any configuration and geometry. We treat here only the problem inspired by the actin comet formation on rigid beads. \section{Results} Having defined the mechanical and dynamical equations in a consistent manner, we will now explore some consequences. We first consider the case of an axisymmetric state of the network and analyse then its linear stability with respect to a morphological perturbation. \subsection{Axisymmetric network} In the axisymmetric state the network geometry is described by the "radial layer number" $\Lambda_1$. Recall that if in the discrete network the true number of nodes in the radial direction is $N_1$ (with $N_1$ being a large integer), in the continuum limit the relation $\Lambda_1= N_1 \eta=N_1l_0/R_0$ holds. We look for an axisymmetric configuration of the form $\boldsymbol{ \phi}=\phi_R \mathbf{\hat r}.$ Equations (\ref{en}-\ref{bc2}) are solved numerically using continuation methods \cite{AUTO97}. Figure \ref{hdmu} (a) shows the observable thickness $h$ of the network $h=\phi_R(\Lambda_1)-R_0$ in the mechanical equilibrium configuration depending on the radial layer number $\Lambda_1$. The distinction between the observable thickness and the radial layer number is important, since they are not necessarily related in a linear manner. \begin{figure}[hbt] \includegraphics[width=0.8\hsize]{figure2.eps} \caption{Observable network thickness $h$ depending on the radial layer number $\Lambda_1$.\mylab{hdmu}} \end{figure} For thin networks $h$ increases linearly with $\Lambda_1$ and one finds $h/R_0=\sqrt{3}\Lambda_1/2$. For larger networks the tangential tension at the outer network interface leads to a radial compression and $h$ nearly saturates for $\Lambda_1 \simeq 0.8$. An interesting property follows naturally from our formulation, which is that for an axisymmetric state the elastic chemical potential difference is identical at the two interfaces, denoted as $\Delta\mu^{(0)}_m$. If the thickness of the network is modified, there is no way to discriminate between the fact that material has been exchanged at the cylinder surface or at the free surface, only the initial and the final thickness matter, and not the path followed by the system. This property is not only comforting the theory, but can even be exploited to extract some interesting results. Indeed, it is possible to determine analytically the chemical potential at the external surface, which is linked to the observable network thickness $h$ in a simple way (see Appendix \ref{der:pot} for details) \begin{equation} \Delta\mu_m^{(0)}={kh^2\over 8 l_0 R_0^2}\left(h+2R_0\right)^2\,, \mylab{mum} \end{equation} where we have used the fact that the filaments of type 1 and 3 (cf. Fig. \ref{network:topo}) are at equilibrium length [to ensure a vanishing traction force (\ref{bc2})] and the filaments of type 2 have length $l_2=l_0/R_0\phi_R(\Lambda_1)=l_0(h+R_0)/R_0$. Now we consider growth of an axisymmetric network whereby polymerization only occurs at the internal interface with $\Delta\mu_c<0$. Dendritic actin networks nucleated by the Arp2/3 complex typically show a kinetic polarity with a rapidly polymerizing 'plus'-end interface directed towards the nucleating surface i.e. $\Delta\mu_{ci}<0$, and a slowly depolymerizing 'minus'-end oriented towards the solvent, i.e. $\Delta\mu_{ce}>0$ \cite{PBM00}. For more clarity we consider here only the case, that the motility medium does not contain any fragmentation proteins, such as cofilin and we treat the limiting case that growth only occurs at the internal interface while the external interface is stationary. The model can be straightforwardly extended to two dynamical interfaces, which gives for example the treadmilling behavior. If growth occurs only at the internal interface the interface velocity (\ref{vn}) vanishes in steady state and one finds in the limit $h\ll 2R_0$ for the observable stationaray network thickness \begin{equation} \bar h = \sqrt{-{2l_0 \Delta\mu_c \over k}}, \mylab{hss} \end{equation} where $\bar h$ stands for the steady thickness. Recall that in steady state the radial layer number is related to the observable thickness by $\bar h=\phi_R(\bar\Lambda_1)-R_0$. Expression (\ref{hss}) can be further related to the network properties and geometry and to the known rate equations of actin polymerization, which offers interesting scaling relations, which will be discussed at the end of this Letter. The next important step is to analyze the linear stability against modulations of the gel thickness. \subsection{Linear stability analysis} We introduce a small perturbation of amplitude $\varepsilon$ at the internal interface, while the structure of the external interface remains unchanged. The perturbation is encoded in the growth velocity at the internal surface. Since we consider a system, which is periodic (period $2\pi$) and translationally invariant in $\lambda_2$ a perturbation of any quantity (shape, strain, growth velocity) can be expressed in terms of a $\cos$-series with wavenumber $q$ and growth rate $\beta_q$. We have solved the model equations in the linear regime. \begin{figure}[htb] \includegraphics[width=0.8\hsize]{figure3.eps} \caption{Dispersion relation for the shear-free (slip BC) and the Dirichlet (fixed BC) boundary condition. Parameters are $\bar\Lambda_1=0.7$. The time scale is $t_0=l_0/(MkR^2_0)$. The inset shows the positive growth rate $\beta_1$ depending on the radial layer number $\Lambda_1$.\mylab{disprel}} \end{figure} Figure \ref{disprel} shows a typical dispersion relation. Also shown is the dispersion relation for an alternative boundary condition, that the network is fixed and cannot slide on the cylinder surface, i.e. condition (\ref{bc1}) is replaced by $\boldsymbol{\phi}=R_0\mathbf{ \hat r}$ at the internal interface. A robust and interesting outcome is that only perturbations with the wave number $q=1$ are unstable (i.e. $\beta_1>0$), while all other modes are stable. For relatively thin networks we find $\beta_1 = MkR_0^2 \bar\Lambda_1^2/ l_0$. This instability is only present when the network can slide on the cylinder surface. When the bonds are fixed, no instability is found. The mode selection arises naturally within the model. Damping of high modes is naturally present in the model due to the accumulation of elastic shear-stresses in the bulk. The instability corresponding to wave number $q=1$ means, that the network shrinks at one side of the cylinder and grows at the opposite side of the cylinder, i.e. the instability initiates the formation of an actin comet as observed in biomimetic motility experiments\cite{OuT99,GPP05,AMMG10,UCA03,GFT03,DSR08,BCJ04}. \begin{figure} \includegraphics[width=0.45\hsize]{figure4a.eps} {\Large a} \includegraphics[width=0.45\hsize]{figure4b.eps}{\Large b} \caption{Network shape before (a) and after symmetry-breaking (b). (b) was obtained by superimposing the axisymmetric shape and the unstable mode with the amplitude $0.15\,\Lambda_1$. The color encodes the spatial distribution of the strain energy density (a) and the change in strain energy density (b). The energy density scale is $k/l_0$.\mylab{comet}} \end{figure} Figure \ref{comet} shows typical shapes of the network before and after symmetry-breaking. Also shown is the strain energy density [Fig. \ref{comet} (a)] and the change in strain energy density compared to the axisymmetric shape [Fig. \ref{comet} (b)]. Interestingly, the symmetry is broken by increasing the strain energy density in the thin network regions (front) and by decreasing it in the thick network regions (back). As time evolves the instability should be amplified and other mechanisms (e.g. fracture) may get important and lead to a larger comet. A detailed study of the subsequent nonlinear regime will be dealt with in the future. \subsection{Scaling properties and comparison with experiments} { In the following we will discuss some scaling relations (see Appendix \ref{scaling} for more details). We base our calculations on the assumptions, that the filaments behave as entropic springs with persistance length $l_p$ \cite{GSMM04}, and that free monomers (concentration $c$, size $l_m$) polymerize into linear filaments with the known kinetics $k_+(c-c_c)$ in the absence of mechanical stress \cite{PBM00}, where $k_+$ denotes the rate constant of polymerization and $c_c$ denotes the critical monomer concentration, at which the polymerization speed is zero. We find the following scaling for the observable stationary network thickness (\ref{hss}) \begin{equation} (\bar h/R_0)^2\sim {l_0^3(c-c_c)\over l_p^2l_m c_c} \end{equation} Using typical values for $l_0=100\,$nm, $l_p=10\,\mu$m, $l_m=3\,$nm and $c/c_c=10$ we find ${\bar h/ R_0}\sim 0.2$. This value and the linear scaling of $\bar h$ with $R_0$ are consistent with experiments \cite{GPP05}. The growth rate scales as $\beta\sim k_+(c-c_c)l_m/R_0$ which yields a time scale for the birth of the instability \begin{equation} \tau \sim {R_0\over l_m k_+(c-c_c)} \end{equation} Using $R_0=1\,\mu$m and $k_+(c-c_c)=1\,$s$^{-1}$ \cite{PBM00} one finds a typical time $\tau=5\,$min, which is a reasonable value \cite{GPP05}. \section{Discussion} We have provided a general theoretical framework for stress and actin growth coupling. Application to actin growth on beads led to the following major results: (i) the surprising mode selection $q=1$ and the stabilization of all higher modes without needing an ad-hoc cut-off length, (ii) the role of the boundary conditions at the nucleating surface (sliding/fixed) for the existence of the instability, (iii) the scaling laws which are consistent with experiments. The instability reported here is induced by growth and not by fracture. Experiments on vesicles \cite{DSR08} as a nucleating surface are consistent with this mechanism. They suggest, that for vesicles symmetry-breaking does not occur via a fracture mechanism at the external network interface, but via a variation in the growth speed along the internal vesicle/network interface. In \cite{DSR08} it was observed that shortly after symmetry-breaking the newly formed network layers at the internal interface had a varying thickness (thin on one side, thick on the opposite side of the vesicle), wereas older network layers at the external network interface maintained a homogeneous thickness. This observation is consistent with an asymmetric network polymerization at the internal interface. It is reasonable to expect bonds at the membrane to slide due to the fluid nature of the membrane, supporting our outcome that only in this case an instability takes place. For experiments with rigid beads \cite{GPP05} it is not obvious (albeit not excluded) that bonds can slide, but still an instability could be observed. It has been proposed by the authors \cite{GPP05} that the instability is triggered by fracture of the network. In preliminary calculations we have also identified a morphological instability of purely mechanical origin. However, this instability requires a critical network thickness, beyond the thickness which we considered here. A further investigation of this problem as well as other growth geometries (lamellipodium) and the effect of active stresses \cite{KJJP04} will be part of future work. \acknowledgments C.M. acknowledges financial support from CNES (Centre d'Etudes Spatiales). The LIPhy is part of the LabEx Tec 21 (Investissements de l'Avenir - grant agreement no. ANR-11-LABX-0030). \begin{appendix} \section{Derivation of the elastic chemical potential difference} \mylab{der:pot} In the derivation of the elastic chemcical potential we will assume a strain energy functional with the boundary conditions defined in Eqs. (\ref{en}-\ref{bc2}) in the main text. The elastic chemical potential difference $\Delta\mu_m$ for inserting a material element at the network interface can be calculated from the variation of the elastic strain energy $F$ with respect to a change in shape of the elastic body in the material frame $\Omega$ by respecting the prescribed boundary conditions \begin{equation} \Delta\mu_m={\delta F\over \delta\Omega}\,. \end{equation} Upon a modification of $\Omega$ by $\delta\Omega$ the change in the strain energy is given to lowest order in $\delta\Omega$ by \begin{equation} \delta F={1\over \eta^2}\int_\Omega \mathbf{T}_i \cdot\delta\mathbf{h}_i \,d\lambda_1d\lambda_2 +{1\over \eta^2} \int_{\delta\Omega} f d\lambda_1d\lambda_2 \mylab{deltaF} \end{equation} with $\mathbf{T}_i=\partial f/\partial \mathbf{h}_i$ and $\mathbf{ h}_i=\eta \partial_{\lambda_i}\boldsymbol{\phi}$. Upon partial integration of (\ref{deltaF}) one finds \begin{eqnarray} \delta F & = & -{1\over \eta} \int_\Omega \partial_{\lambda_i} \mathbf{T}_i\cdot\delta\boldsymbol{\phi}\, d\lambda_1d\,\lambda_2 +{1\over \eta}\int_S \mathbf{T}\cdot\delta\boldsymbol{\phi}\, ds + \nonumber\\ & & {1\over \eta^2} \int_{\delta\Omega} f d\lambda_1d\lambda_2 \mylab{deltaF2} \end{eqnarray} with $\mathbf{T}=\mathbf{T}_1\nu_1+\mathbf{T}_2\nu_2$ and where $\nu_1$ and $\nu_2$ denote the components of the unit outward normal vector $\boldsymbol{\nu}$ in the material frame $(\lambda_1,\lambda_2)$. The first integral in (\ref{deltaF2}) vanishes due to the mechanical equilibrium condition $\partial_{\lambda_i}\mathbf{T_i}=0$ and only the second and third integral in (\ref{deltaF2}) will contribute to the chemical potential difference. Now, we assume that the perturbation of the interface is described by a function $\epsilon(|s-s_0|)\boldsymbol{\nu}(s)$, with $\epsilon(|s-s_0|)\ll 1$. Since the perturbation of the boundary is local and $\epsilon$ is decaying rapidly with $|s-s_0|$ we can evaluate the third integral in (\ref{deltaF2}) to the lowest order \begin{equation} \int_{\delta\Omega} f \,d\lambda_1\,d\lambda_2=\int_S f(s)\epsilon(|s-s_0|)\,ds=f(s_0)\delta\Omega\mylab{pert:prop} \end{equation} The second integral over the boundary of the unperturbed domain depends crucially on the boundary condition. First we will consider the simple case of a material exchange at the external force-free interface with $\mathbf{T}=0$. In this case also the line integral in (\ref{deltaF2}) vanishes and we find \begin{equation} \delta F={1\over \eta^2}f(s_0)\delta\Omega, \end{equation} which leads to the elastic chemical potential difference for a perturbation at $s_0$ at the external interface \begin{equation} \Delta\mu_{me}={1\over \eta^2}f(s_0) \mylab{dmue} \end{equation} In the axisymmetric configuration (\ref{dmue}) can be linked to the observable network thickness $h=\phi_R(\Lambda_1)-R_0$ in a simple way. Making use of the fact that the traction vector $\mathbf{T}$ at the external interface vanishes for $l_1=l_3=l_0$ and since $l_2=\eta\phi(\Lambda_1)=\eta(R_0+h)$, one finds after some manipulation of (\ref{dmue}) one recovers Eq. (\ref{mum}) from the main text \begin{equation} \Delta\mu_{me}^{(0)}={kh^2\over 8 l_0 R_0^2}\left(h+2R_0\right)^2. \end{equation} Next we consider the more complex case of a material exchange at the internal interface. Taking advantage of the boundary conditions [Eq. (\ref{bc1}) in the main text] we note first that after a modification of the internal boundary we find for the position of the internal boundary up to lowest order in the perturbation \begin{equation} R_0^2=\boldsymbol{\phi}\cdot\boldsymbol{\phi}+2\boldsymbol{\phi}\cdot(\delta\boldsymbol{\phi}+\partial_{\lambda_i}\boldsymbol{\phi}\nu_i\epsilon). \end{equation} The position of the boundary before the perturbation $\boldsymbol{\phi}$ fulfills $|\boldsymbol\phi|=R_0$ and it follows that \begin{equation} 0=\boldsymbol{\phi}\cdot(\delta\boldsymbol{\phi}+\partial_{\lambda_i}\boldsymbol{\phi}\nu_i\epsilon)\quad \mbox{and}\quad\boldsymbol{\phi}\cdot\delta\boldsymbol{\phi}=-\boldsymbol{\phi}\cdot\partial_{\lambda_i}\boldsymbol{\phi}\nu_i\epsilon.\mylab{trans} \end{equation} Since the vectors $\boldsymbol{\phi}$ and $\mathbf{T}$ are parallel to the normal direction of the boundary we can write using identity (\ref{trans}) \begin{equation} \mathbf{T}\cdot\delta\boldsymbol{\phi}={1\over R_0^2}(\mathbf{T}\cdot \boldsymbol{\phi})(\boldsymbol{\phi}\cdot\delta\boldsymbol{\phi})=-\mathbf{T}\cdot \partial_{\lambda_i}\boldsymbol{\phi}\nu_i\epsilon. \end{equation} After an elementary manipulation we find from (\ref{deltaF2}) and (\ref{pert:prop}) \begin{eqnarray} \delta F & = & -{1\over \eta^2}\int_S \mathbf{T}\cdot \mathbf{h}\ \epsilon (|s-s_0|)\,ds +{1\over \eta^2} f(s_0)\delta\Omega\nonumber\\ & = & {1\over \eta^2} \left[f-\mathbf{T}\cdot \mathbf{h}\right]_0\delta\Omega \mylab{dFi}, \end{eqnarray} with $\mathbf{h}=\eta\partial_{\lambda_i}\boldsymbol{\phi}\nu_i$ and where the subscript $_0$ indicates that all quantities are calculated at the position $s_0$. From (\ref{dFi}) we infer the elastic chemical potential difference at the internal interface [Eq. (\ref{mumi}) in the main text] \begin{equation} \Delta\mu_{mi}={1\over \eta^2} \left[f-\mathbf{T}\cdot\mathbf{h}\right]. \end{equation} \section{Scaling behavior} \mylab{scaling} Here we demonstrate some scaling relations, relevant for biomimetic experiments on actin driven motility. First we note, that the polymerization part of the chemical potential $\Delta\mu_c$ used in Eq. (\ref{mumi}) in the main text is based on a material exchange in Lagrangian coordinates. Therefore, the change in chemical potential per node is given by $\eta^2\Delta\mu_c$, which can be related to the chemical potential difference per monomer $\Delta\tilde\mu_c$ by assuming that each node contains 3 filaments of length $l_0$ and that each filament contains $l_0/l_m$ monomers of size $l_m$. Consequently \begin{equation} \Delta\tilde\mu_c={l_0l_m\over 3R_0^2}\Delta\mu_c=-k_BT\ln{c\over c_c} \mylab{muc} \end{equation} where $c$ denotes the actual concentration of monomers in solution, $c_c$ denotes the critical monomer concentration where the polymerization speed vanishes and $k_BT$ denotes the thermal energy. Using expression (\ref{muc}) we can write \begin{equation} \left({\bar h\over R_0}\right)^2=-{2l_0\Delta\mu_c\over k R_0^2}={6k_BT\over k l_m}\ln{c\over c_c} \end{equation} Assuming now, that the filaments behave as entropic springs with spring constant $k=k_BTl_p^2/l_0^3$ \cite{GSMM04}, where $l_p$ denotes the persistence length we find \begin{equation} \left({\bar h \over R_0}\right)^2= {6l_0^3 \over l_p^2l_m}\ln{c\over c_c} \approx {6l_0^3 \over l_p^2l_m c_c} (c-c_c) \mylab{hR0} \end{equation} Note the linear scaling of $\bar h$ with $R_0$ as observed experimentally in \cite{GPP05}. Using typical values for $l_0=100\,$nm, $l_p=10\,\mu$m, $l_m=3\,$nm and $c/c_c=10$ one finds $\bar{h}/R_0\approx 0.2$. Now, we use the known kinetic eqation of polyerization of actin to estimate the time scale of symmetry breaking $t_0=l_0/(MkR_0^2)$, i.e. we estimate the mobility constant $M$. Typically the polymerization of actin without mechanical stresses is described by the following kinetic equation \cite{PBM00} \begin{equation} v_p=k_+(c-c_c)\mylab{vp} \end{equation} where $k_+$ denotes the rate constant of polymerization. The normal interface velocity [Eq. (\ref{vn}) in the main text] is related to the polymerization speed by $v_n\approx l_m/R_0 v_p$ and we find \begin{eqnarray} v_p & = & {R_0\over l_m}v_n=-{R_0\over l_m}M\Delta\mu_c={3MR_0^3k_BT\over l_0l_m^2}\ln{c\over c_c} \nonumber\\ & = &{3MR_0^3k_BT\over l_0l_m^2c_c} (c-c_c) \mylab{vpc} \end{eqnarray} Comparisson of (\ref{vpc}) with (\ref{vp}) gives the following expression for the mobility constant \begin{equation} M={k_+c_cl_0l_m^2 \over 3R_0^3 k_BT} \end{equation} and the time scale $t_0$ can be rewritten as \begin{equation} t_0={l_0\over M k R_0^2}={3R_0l_0^3\over k_+c_c l_p^2l_m^2} \mylab{t0} \end{equation} As stated in the main text, the growth rate for symmetry-breaking scales as $\beta_1\sim \bar\Lambda^2/t_0$ for relatively thin networks. Assuming $\bar\Lambda_1^2\sim (\bar h/R_0)^2$ and using (\ref{hR0}) and (\ref{t0}) we find the following scaling $\beta_1\sim k_+(c-c_c)l_m/R_0$. Consequently the typical time scale for symmetry-breaking scales linearly with $R_0$ as shown experimentally in \cite{GPP05}. For $R_0=1\,\mu$m and $k_+(c-c_c)=1$\,$s^{-1}$ \cite{PBM00} one finds a typical time scale of symmetry-breaking of 5\,min, which is a reasonable result \cite{GPP05}. \end{appendix}
\section{Introduction} The research of cycles has been fundamental since the beginning of graph theory. One of various problems on cycles which have been considered is the study of cycle lengths modulo a positive integer $k$. Burr and Erd\H{o}s \cite{Erd76} conjectured that for every odd $k$, there exists a constanct $c_k$ such that every graph with average degree at least $c_k$ contains cycles of all lengths modulo $k$. In \cite{B77}, Bollob\'as resolved this conjecture by showing that $c_k= 2[(k+1)^k-1]/k$ suffices. Thomassen \cite{Th83,Th88} strengthened the result of Bollob\'as by proving that for every $k$ (not necessarily odd), every graph with minimum degree at least $4k(k+1)$ contains cycles of all even lengths modulo $k$, which was improved to the bound $2k-1$ by Diwan \cite{D10}. Note that in case $k$ is even, any integer congruent to $l$ modulo $k$ has the same parity with $l$, and thus we can not expect that there are cycles of all lengths modulo $k$ in bipartite graphs (even with sufficient large minimum degree). On the other hand, Thomassen \cite{Th83} showed that for every $k$ there exists a least natural number $f(k)$ such that every non-bipartite 2-connected graph with minimum degree at least $f(k)$ contains cycles of all length modulo $k$. Here the 2-connectivity condition can not be further improved, as one can easily construct a non-bipartite connected graph with arbitrary large minimum degree but containing a unique (also arbitrary) odd cycle. Thomassen \cite{Th83} remarked that the upper bound for $f(k)$ obtained by him is perhaps ``far too large'' and asked if $f(k)$ can be bounded above by a polynomial. Bondy and Vince \cite{BV} resolved a conjecture of Erd\H{o}s by showing that every graph with minimum degree at least 3 contains two cycles whose lengths differ by one or two. Verstra\"ete \cite{V00} proved that if graph $G$ has average degree at least $8k$ and even girth $g$ then there are $(g/2-1)k$ cycles of consecutive even lengths in $G$. In an attempt to extend the result of Bondy and Vince, Fan \cite{Fan02} showed that every graph with minimum degree at least $3k-2$ contains $k$ cycles of consecutive even lengths or consecutive odd lengths. Sudakov and Verstra\'ete proved \cite{SV08} that if graph $G$ has average degree $192(k+1)$ and girth $g$ then there are $k^{\flo{(g-1)/2}}$ cycles of consecutive even lengths in $G$, strengthening the above results in the case that $k$ and $g$ are large. It is natural to ask if one can pursue the analogous result for odd cycles. In this paper, we show that this indeed is the case by the following theorem. \begin{thm}\label{thm:main} There exists an absolute constant $c>0$ such that for every natural number $k$, every non-bipartite 2-connected graph $G$ with average degree at least $ck$ and girth $g$ contains at least $k^{\flo{(g-1)/2}}$ cycles of consecutive odd lengths. \end{thm} \noindent We point out that the non-bipartite condition here is necessary and the 2-connectivity conditions can not be improved. In view of the {\it Moore Bound}, our lower bound on the number of cycles is best possible (up to constant factor) for infinitely many integers $k$ when $g\le 8$ or $g=12$. And more generally, the well-known conjecture that the minimal order of graphs with minimal degree $k$ and girth $g$ is $O(k^{\flo{(g-1)/2}})$ would imply that our result gives the correct order of magnitude for other values of $g$. Let $G$ be a graph as in Theorem \ref{thm:main}. It is clear that there are at least $k$ cycles of consecutive odd lengths in $G$, which assures that $G$ contains cycles of all odd lengths modulo $k$ (whenever $k$ is even or odd). Together with the aforementioned result of Diwan on cycles of all even lengths modulo $k$, we answer the question of Thomassen by improving the upper bound of $f(k)$ to a linear function by the following corollary. \begin{coro}\label{coro:cycle_mod_k} There exists an absolute constant $d>0$ such that for every natural number $k$, every non-bipartite 2-connected graph with minimum degree at least $dk$ contains cycles of all lengths modulo $k$. \end{coro} \noindent This bound is sharp up to the constant factor: the complete graph $K_{k+1}$ contains cycles of all lengths but $2$ modulo $k$ and thus shows that $f(k)\ge k+1$. \medskip All graphs considered are simple and finite. Let $G$ be a graph. We denote the number of vertices in $G$ by $|G|$, the vertex set by $V(G)$, the edge set by $E(G)$ and the minimum degree by $\delta(G)$. If $S\subset V(G)$, then $G-S$ denotes the subgraph of $G$ obtained by deleting all vertices in $S$ (and all edges incident with some vertex in $S$). If $S\subset E(G)$, then $G-S$ is obtained from $G$ by deleting all edges in $S$. Let $A$ and $B$ be subsets of $V(G)$. An {\it $(A,B)$-path} in $G$ is a path with one endpoint in $A$ and the other in $B$. If $A$ only contains one vertex $a$, then we simply write $(A,B)$-path as $(a,B)$-path. We say a path $P$ is {\it internally disjoint} from $A$, if no vertex except the endpoints in $P$ is contained in $A$. The rest of paper is organized as follows. In next section, we establish Theorem \ref{thm:main} based on the approach of \cite{SV08}. The last section contains some remakes and open problems. We make no effort to optimize the constants in proofs and instead aim for simpler presentation. \section{The proof} Before processing, we state the following functional lemma from \cite{V00}, which will be applied multiple times and become essential in the proof of our main theorem. \begin{lem} {\rm (Verstra\"ete \cite{V00})} \label{lem:A-Bpaths} Let $C$ be a cycle with a chord, and let $(A,B)$ be a nontrivial partition of $V(C)$. Then $C$ contains $(A,B)$-paths of every length less than $|C|$, unless $C$ is bipartite with bipartition $(A,B)$. \end{lem} \begin{proof}[Proof of Theorem \ref{thm:main}] We shall show that it suffices to use $c=456$. Let $G$ be a non-bipartite 2-connected graph with average degree at least $456\cdot k$ and girth $g$. Our goal is to show that $G$ contains $t:=k^{\flo{(g-1)/2}}$ cycles of consecutive odd lengths. As it holds trivially when $k=1$, we assume that $k\ge 2$. Let $G_b$ be a bipartite subgraph of $G$ with the maximum number of edges. It is easy to see that $G_b$ is a connected spanning subgraph of $G$ with average degree at least $228k$. Since $G$ is non-bipartite, there exists an edge $xy\in E(G)$ such that both $x$ and $y$ lie in the same part of the bipartition of $G_b$. Let $T$ be the breadth first search tree in $G_b$ with root $x$, and let $L_i$ be the set of vertices of $T$ at distance $i$ from its root $x$ for $i\ge 0$ (so $L_0:=\{x\}$). As $T$ is also a spanning tree, it follows that $V(G)=V(G_b)=\cup_{i\ge 0} L_i$. For any two vertices $a$ and $b$ in the tree $T$, we denote $T_{ab}$ by the unique path in $T$ with endpoints $a$ and $b$. By the choice of edge $xy$, clearly the vertex $y$ lies in $L_{2l}$ for some integer $l\ge 1$, therefore the cycle $D:=T_{xy}\cup xy$ is an odd cycle in $G$. By the definition of $T$, every edge of $G_b$ joints one vertex in $L_i$ to the other in $L_{i+1}$ for some $i\ge 0$. Thus, we have $$\sum_{i} e(L_i,L_{i+1})=e(G_b)\ge 114k\cdot |G_b|=114k\cdot \sum_i |L_i|\ge 57k\cdot \sum_{i}\left(|L_i|+|L_{i+1}|\right).$$ So there must exist some $i\ge 0$ such that the induced (bipartite) subgraph $G_i:=G_b[L_i\cup L_{i+1}]$ has average degree at least $114k$. We now use the following lemma in \cite{SV} to find a long cycle with at least one chord (in fact with many chords) in $G_i$. \vskip 2mm \begin{lem} {\rm (Sudakov and Verstra\"ete \cite{SV})} Let $G$ be a graph of average degree at least $12(d+1)$ and girth $g$. Then $G$ contains a cycle $C$ with at least $\frac{1}{3}d^{\flo{(g-1)/2}}$ vertices of degree at least $6(d+1)$, each of which has no neighbors in $G-V(C)$. \end{lem} \vskip 2mm Note that $t=k^{\flo{(g-1)/2}}$. Since $114k\ge 12(9k+1)$ for $k\ge 2$ and the girth of $G_i$ is at least $g$, this lemma shows that there is a cycle $C$ with a chord in $G_i$, satisfying \begin{equation}\label{equ:|C|} |C|\ge \frac{1}{3}\cdot (9k)^{\flo{(g-1)/2}}\ge 2(t+1). \end{equation} We notice that $(V(C)\cap L_i, V(C)\cap L_{i+1})$ is the unique bipartition of $C$ (this is for the use of Lemma \ref{lem:A-Bpaths} later). Let $T'$ be the minimal subtree of $T$ whose set of leaves is precisely $V(C)\cap L_i$, and let $z$ be the root of $T'$ (i.e., the one at the shortest distance from $x$). By the minimality of $T'$, $z$ has at least two branches of $T'$. Let the depth of $T'$ (i.e., the distance between its root $z$ and its leaves) be $j$. Recall that $D$ is an odd cycle in $G$ consisting of $T_{xy}$ and edge $xy$, where $y\in L_{2l}$. Depending on whether $V(D)$ interests $V(C)\cup V(T')-\{z\}$ or not, we distinguish the following two cases. \bigskip {\bf Case 1.} $V(D)$ and $V(C)\cup V(T')-\{z\}$ are disjoint. \vskip 1mm In this case, the tree $T$ contains a path $Z$ from $z$ to $V(D)$, which is internally disjoint from $V(C)\cup V(T')$. Note that $G$ is 2-connected, so there are two disjoint $(V(C)\cup V(T'), V(D))$-paths, say $P$ and $Q$, in $G$, which are internally disjoint from $V(C)\cup V(T')\cup V(D)$. Routing $P,Q$ through $Z$ if necessary, we may assume that $P$ is from $z$ to $p\in V(D)$, and $Q$ is from $w\in V(C)\cup V(T')-\{z\}$ to $q\in V(D)-\{p\}$. Base on the location of $w$, we divide the remainder of this case into two parts. Let us first consider when $w\in V(T')-\{z\}$. Let $A$ be the set of all leaves in the subtree of $T'$ with root $w$, and let $B:=V(C)-A$. As $w\neq z$, we see $B^*:=V(C)\cap L_i-A$ is nonempty, which shows that $(A,B)$ is not the bipartition of $C$. By Lemma \ref{lem:A-Bpaths} and the equation \eqref{equ:|C|}, $C$ contains $(A,B)$-paths of all even lengths up to $2t+1$, all of which in fact are $(A,B^*)$-paths because $C$ is bipartite. To find $t$ cycles of consecutive odd lengths, now it is enough to show that for any $a\in A$ and $b\in B^*$, there exists an $(a,b)$-path in $G-E(C)$ with a fixed odd length. To see this, first note that paths $T'_{aw}$ and $T'_{bz}$ are disjoint and of fixed lengths; since $D$ is an odd cycle, one can choose a $(p,q)$-path $R$ in $D$ (out of two choices) such that the $(a,b)$-path $T'_{aw}\cup Q\cup R\cup P\cup T'_{zb}$ in $G-E(C)$ is of a fixed odd length. Now we consider the situation when $w\in V(C)-V(T')$. So $w\in V(C)\cap L_{i+1}$, and clearly $(\{w\},V(C)-\{w\})$ is not a bipartition of $C$. By Lemma \ref{lem:A-Bpaths} and the equation \eqref{equ:|C|}, $C$ contains $(w,V(C)-\{w\})$-paths of all odd lengths up to $2t+1$, all of which are $(w, V(C)\cap L_i)$-paths because $C$ is bipartite. For any $u\in V(C)\cap L_i$, the length of path $T'_{uz}$ is fixed, that is the depth $j$ of $T'$. Therefore, $C\cup T'$ contains $(z,w)$-paths of all lengths in $\{1+j,3+j,\ldots,2t+1+j\}$. Since $D$ is an odd cycle, similarly as in the last paragraph one can choose a $(p,q)$-path $R$ in $D$ such that the $(z,w)$-path $P\cup R\cup Q$ in $G-E(C\cup T')$ is of the same parity as $j$. Putting the above paths together, we see that $G$ contains at least $t$ cycles of consecutive odd lengths. This completes the proof of Case 1. \bigskip {\bf Case 2.} $V(D)$ intersects $V(C)\cup V(T')-\{z\}$. \vskip 1mm Let $w\in V(D)\cap \left(V(C)\cup V(T')-\{z\}\right)$ be the vertex such that $T_{wy}$ is the shortest path among all choices of $w$. Note that now $T_{wz}=T'_{wz}$ is a subpath in the cycle $D$. Again we distinguish on the location of $w$. First we consider when $w\in V(T')-\{z\}$. Let $A$ be the set of all leaves in the subtree of $T'$ with root $w$, and $B^*:=V(C)\cap L_i-A$. By the same proof as in the second paragraph of Case 1, we conclude that $C$ contains $(A,B^*)$-paths of all even lengths up to $2t+1$. We also notice that $D$ consists of two $(w,z)$-paths $T'_{wz}$ and $P:=D-T'_{wz}$, whose lengths are of opposite parities. For any $a\in A$ and $b\in B^*$, $T'_{aw}\cup T'_{wz}\cup T'_{zb}$ forms an $(a,b)$-walk with a fixed even length, that is twice of the depth of $T'$. This suggests that $T'_{aw}\cup P\cup T'_{zb}$ is an $(a,b)$-path in $G-E(C)$ with a fixed odd length, which, combining with these $(A,B^*)$-paths with lengths $2,4,\ldots, 2t$ in $C$, comprise $t$ cycles of consecutive odd lengths in $G$. We are left with the case when $w\in V(C)-V(T')$. This shows that $w\in V(C)\cap L_{i+1}$ and $(\{w\}, V(C)-\{w\})$ is not the bipartition of $C$. By Lemma \ref{lem:A-Bpaths} as well as the equation \eqref{equ:|C|}, $C$ contains $(w,V(C)-\{w\})$-paths of all odd lengths up to $2t+1$. Similarly as the previous proofs, these odd paths are actually $(w,V(C)\cap L_i)$-paths. Recall that the depth of $T'$ is $j$. Therefore, $C\cup T'$ contains $(w,z)$-paths of all lengths in $\{1+j,3+j,\ldots, 2t+1+j\}$ and particularly the sub-path $T_{wz}$ of $D$ has length $1+j$. We know $D$ is an odd cycle, so $D-T_{wz}$ is a $(w,z)$-path in $G-E(C\cup T')$ whose length is of the same parity as $j$. Putting $D-T_{wz}$ and these $(w,z)$-paths in $C\cup T'$ together, we find at least $t$ cycles of consecutive odd lengths. This proves Case 2, finishing the proof of Theorem \ref{thm:main}. \end{proof} \section{Concluding remarks} In \cite{BV} Bondy and Vince gave an infinite family of non-bipartite 2-connected with arbitrary large minimum degree but containing no two cycles whose lengths differ by one. This tells that Theorem \ref{thm:main} is sharp from another point of view. The situation changes completely when the connectivity increases. Fan \cite{Fan02} showed that every non-bipartite 3-connected graph with minimum degree at least $3k$ contains $2k$ cycles of consecutive lengths. A conjecture of Dean (see \cite{DLS93}) also considered the connectivity and asserted that every $k$-connected graph contains a cycle of length $0$ modulo $k$. We observe that this is best possible for odd $k$ (if true), as the complete bipartite graph $K_{k-1,k-1}$ is $(k-1)$-connected but has no cycle of length $0$ modulo $k$. Thomassen showed in \cite{Th83-girth} that graphs of minimum degree at least 3 and large girth share many properties with graphs of large minimum degree. For example, he proved that if $G$ is a graph of minimum degree at least 3 and girth at least $2(k^2 + 1)(3\cdot 2^{k^2+1} + (k^2 + 1)^2-1)$, then $G$ contains cycles of all even lengths modulo $k$ (while we have seen the analogous result for graphs of large minimum degree in \cite{B77,Th83,Th88,D10}). We have also seen that graphs of large minimum degree contain cycles of consecutive even lengths (e.g., results from \cite{SV08,V00}), however to the best of our knowledge it is not known if there exists a natural number $g(k)$ such that every graph of minimum degree at least 3 and girth at least $g(k)$ contains $k$ cycles of consecutive even lengths. Similar question can be raised with respect to Theorem \ref{thm:main} as well. In 1970s Erd\H{o}s and Simonovits \cite{ES} asked to determine the {\it chromatic profile} $$ \delta_\chi(H,k):=\inf \{c: \delta(G)\ge c|G| \mbox{ and } H\not\subset G \Rightarrow \chi(G)\le k\}$$ for every graph $H$ (we refer interested readers to \cite{ABGKM} for related topics). Since then, very little is known about $\delta_\chi(H,k)$ for graphs $H$ other than $K_3$. Thomassen \cite{Th07} proved that for every $c>0$ and odd integer $l\ge 5$, every $C_l$-free graph $G$ with minimum degree at least $c|G|$ has chromatic number $\chi(G)$ less than $(l+f(2l-8))/c$. We conclude this paper with the following. \begin{lem} For arbitrary fixed odd integer $l\ge 5$, it holds that $$\Theta\left(\frac{1}{(k+1)^{4(l+1)}}\right)\le \delta_\chi(C_l,k)\le \Theta\left(\frac{l}{k}\right).$$ \end{lem} \begin{proof} The upper bound can be obtained easily by combining Thomassen's result \cite{Th07} and Corollary \ref{coro:cycle_mod_k}. We turn to the lower bound. Let $N(g,k)$ be the minimum $|G|$ over all graphs $G$ with girth at least $g$ and chromatic number at least $k$. The proof of \cite{Mar} shows that $N(g,k)\le \Theta(k^{4g+1})$ when $k\ge 144$. It suffices to prove $\delta_\chi(C_l,k)> k/N(l+1,k+1)$. Let $G_0$ be a graph of minimum order with girth at least $l+1$ and chromatic number at least $k+1$, i.e., $|G_0|= N(l+1,k+1)$, then by the minimality it holds that $\chi(G_0)=k+1$ and $\delta(G_0)\ge k$. We then construct a graph $G$ obtained from $G_0$ by replacing every vertex with an independent set of size $t$ and every edge with a complete bipartite graph. Clearly $G$ contains no $C_l$ (in fact there is no odd cycle of length less than $l+1$ in $G$), where $\chi(G)=\chi(G_0)>k$ and $\delta(G)=t\cdot\delta(G_0)\ge \frac{k|G|}{N(l+1,k+1)}$. This completes the proof. \end{proof}
\section{Introduction} Many classical algorithmic problems on graphs can be defined in terms of finding a subgraph that is isomorphic to a certain pattern graph. For example, the polynomial-time solvable problem of finding perfect matchings and the NP-hard \textsc{Hamiltonian Cycle} and \textsc{Clique} problems arise this way. The goal of the paper is to understand which pattern graphs make this problem easy with respect to polynomial-time solvability and polynomial-time preprocessing. Given graphs~$G$ and~$H$, \SubgraphTest asks if~$G$ has a subgraph isomorphic to the pattern~$H$. Observe that, for every fixed pattern graph $H$, \SubgraphTest is polynomial-time solvable, as we can test each of the $|V(G)|^{|V(H)|}$ mappings from the vertices of $H$ to the vertices of $G$, resulting in a polynomial-time algorithm. Therefore, studying the restrictions of \SubgraphTest to fixed $H$ does not allow us to make a distinction between easy and hard patterns. We can get a more useful framework if we restrict \SubgraphTest to a fixed {\em class} of patterns. For every graph class $\F$, let \FSubgraphTest be the special case of the problem where $H$ is restricted to be in $\F$. For example, if $\F$ is the set of all matchings (1-regular graphs), then \FSubgraphTest is the polynomial-time solvable maximum matching problem; if $\F$ is the set of all cliques, then \FSubgraphTest is the NP-hard \textsc{Clique} problem. Our goal is to understand which classes $\F$ make \FSubgraphTest tractable. We also investigate a well-studied and natural variant of finding subgraphs. Given graphs $G$ and $H$, and an integer $t$, \Packing asks if $G$ has $t$ vertex-disjoint subgraphs isomorphic to $H$. Unlike for \SubgraphTest, now it makes sense to define the problem \HPacking for a fixed graph $H$: for example, $K_2$-\textsc{Packing} is the polynomial-time solvable maximum matching problem and $K_3$-\textsc{Packing} is the NP-hard vertex-disjoint triangle packing problem. We also define the more general \FPacking problem, where $H$ is restricted to be a member of $\F$. \textbf{Kernels and Turing kernels.} Besides looking at the polynomial-time solvability of these problems, we also explore the possibility of efficient preprocessing algorithms, as defined by the notion of polynomial kernelization in parameterized complexity \cite{DowneyF13,FlumG06,LokshtanovMS12}. We can naturally associate a parameter $k$ to each instance measuring the size of the solution we are looking for, that is, we define the parameter $k:=|V(H)|$ for \kSubgraphTest and $k:=t\cdot |V(H)|$ for \kPacking. We say that a problem with parameter $k$ is {\em fixed-parameter tractable (FPT)} if it is solvable in time $f(k)\cdot n^{O(1)}$ for some computable function $f$. The fixed-parameter tractability of various cases of \SubgraphTest is a classical topic of the parameterized complexity literature. It is known that \kFSubgraphTest is FPT if \F is the set of paths \cite{AlonYZ95,Bjorklund10,KneisMRR06,DBLP:journals/siamcomp/WilliamsW13} and, more generally, if \F is a set of graphs of bounded treewidth \cite{AlonYZ95,DBLP:journals/jcss/FominLRSR12}. The case where~\F is the set of all bicliques (complete bipartite graphs), corresponding to the \textsc{Biclique} problem, was a tantalizing open problem for many years. In a recent breakthrough result, Binkai Lin~\cite{Lin14} proved that \textsc{Biclique} is W[1]-hard. In this paper, we study only a specific aspect of fixed-parameter tractability. A {\em polynomial (many-one) kernelization} is a polynomial-time algorithm that creates an equivalent instance whose size is polynomially bounded by the parameter $k$. Intuitively, a kernelization is a preprocessing algorithm that does not solve the problem, but assuming that the parameter value is ``small'' compared to the size of the input, creates a compact equivalent instance by somehow getting rid of irrelevant parts of the input. In the case of \kSubgraphTest, we want to create an equivalent instance with size bounded by $|V(H)|^{O(1)}$: if the pattern $H$ is small compared to $G$, we want to compress the instance to a ``hard core'' that has size comparable to $H$. In recent years, the existence of polynomial kernelization for various parameterized problems has become a thoroughly investigated subject. In 2008, Bodlaender et al.~\cite{BodlaenderDFH09} built on a theorem by Fortnow and Santhanam~\cite{FortnowS11} to introduce the lower bound technology of OR-compositions, which allows us to show that certain parameterized problems do not admit polynomial kernels, unless \containment and the polynomial-time hierarchy collapses to the third level~\cite{Yap83}. In particular, they showed that \kPath (given an undirected graph~$G$ and integer~$k$, does~$G$ contain a simple path of length~$k$?) does not admit a polynomial kernel under this complexity assumption. This work has been followed by a flurry of results refining this technology \cite{BodlaenderJK14,DellM12,DellM10,Drucker12,HermelinW12} and using it to prove negative results for concrete parameterized problems (e.g., \cite{Binkele-RaibleFFLSV12,BodlaenderFLPST09,BodlaenderTY11,CyganKPPW12,DomLS09,FominLMS12,JansenB11,JansenK13,JansenK12,Kratsch13,KratschPRR12}, see also the recent survey of Lokshtanov et al.~\cite{LokshtanovMS12}). We continue this line of research by trying to characterize which \kFSubgraphTest and \kFPacking problems admit polynomial kernels. A natural, but less understood variant of kernelization is Turing kernelization. In a Turing kernelization, instead of creating a single compact instance in polynomial time, we want to solve the instance in polynomial time having access to an oracle solving instances of size $k^{O(1)}$ in constant time. This form of kernelization can be also thought of as some kind of preprocessing: we want to spend polynomial time to preprocess the instance in such a way that the time-consuming part of the work needs to be done on compact instances. While Turing kernelization may seem much more powerful than many-one kernels, there are only a handful of examples where Turing kernelization is possible, but many-one kernelization is not~\cite{AmbalathBHKMPR10,Binkele-RaibleFFLSV12,Jansen14,SchaferKMN12,ThomasseTV13}. On the other hand, the lower bound technology introduced by Fortnow and Santhanam \cite{FortnowS11} and Bodlaender et al.~\cite{BodlaenderDFH09} {\em does not} say anything about the possibility of Turing kernels and therefore we know very little about the limits of Turing kernelization. In fact, even the basic question whether \kPath admits a Turing kernel is open (cf.~\cite{Jansen14}). Hermelin et al.~\cite{HermelinKSWW13} tried to deal with this situation by developing a completeness theory based on certain fundamental satisfiability problems that can be shown to be fixed-parameter tractable by simple branching argument, but for which the existence of polynomial (Turing) kernels is unlikely. They introduced the notion of \textup{WK[1]}-hardness, which can be interpreted as evidence that the problem is unlikely to admit a polynomial Turing kernel.\footnote{It is known~\cite[Lemma 2]{HermelinKSWW13} that the existence of a polynomial-size many-one kernel for a \textup{WK[1]}-hard problem implies \containment.} Unfortunately, Hermelin et al.~\cite{HermelinKSWW13} were unable to prove any hardness result for \kPath; its \textup{WK[1]}-hardness remains an open question. In this paper, we are working under the assumption that \kPath admits no polynomial Turing kernel and interpret the existence of a polynomial-parameter transformation from \kPath to our problem as evidence for the nonexistence of polynomial Turing kernels. Problems for which such a transformation exists will be called \emph{\kPath-hard}. \textbf{Our results.} In this paper, we restrict our study of \FPacking and \FSubgraphTest to hereditary classes $\F$, that is, to classes that are closed under taking induced subgraphs. The polynomial-time solvability of \HPacking is well understood: if every component of $H$ has at most two vertices, then it is a matching problem (hence polynomial-time solvable) and Kirkpatrick and Hell \cite{KirkpatrickH78} proved that \HPacking is NP-hard for every other $H$. It follows that \FPacking is polynomial-time solvable if every component of every graph in $\F$ has at most two vertices, and is NP-hard otherwise. For every fixed $H$, we can formulate \HPacking as a special case of finding $t$ disjoint sets of size $|V(H)|$ each. Hence the problem admits a polynomial kernel of size $t^{O(|V(H)|)}$ using, for example, standard sunflower kernelization arguments~\cite[Appendix A]{DellM12}. However, the exponent of the bound on the kernel size depends on the size of~$H$. Therefore, it does not follow that \FPacking admits a polynomial kernel for every fixed class $\F$, as $\F$ may contain arbitrarily large graphs. Our first result characterizes those hereditary classes $\F$ for which \FPacking admits a polynomial kernel. Interestingly, it seems that Turing kernels are not more powerful for this family of problems: we get the same positive and negative cases with respect to both notions. Let us call a connected bipartite graph {\em $b$-thin} if the smaller partite class has size at most $b$. We say that a graph $H$ is {\em $a$-small/$b$-thin} if every component of $H$ either has at most~$a$ vertices, or is a~$b$-thin bipartite graph (we emphasize that it is possible that $H$ has components of both types). A graph class $\F$ is small/thin if there are $a,b\ge 0$ such that every graph in $\F$ is $a$-small/$b$-thin. \begin{restatable}{\reabctheorem}{restateintropacking}\label{theorem:intro:packing} Let $\F$ be a hereditary class of graphs. If $\F$ is small/thin, then \kFPacking admits a polynomial (many-one) kernel. If $\F$ does not have this property, then \kFPacking admits no polynomial kernel, unless \containment, and moreover it is also \textup{WK[1]}-hard, \textup{W[1]}-hard, or \kPath-hard. \end{restatable} Theorem~\ref{theorem:intro:packing} gives a complete characterization of the hereditary families for which \kFPacking admits a polynomial kernel. It is well known that many problems related to packing small graphs/objects admit polynomial kernels (most of the research is therefore on understanding the exact degree of the polynomial bound \cite{Abu-Khzam10,ChenLSZ07,DellM12,HermelinW12,Moser09}), but we are not aware of any previous result showing that thin bipartite graphs have similar good properties. This revelation about thin bipartite graphs highlights the importance of looking for dichotomy theorems such as Theorem~\ref{theorem:intro:packing}: while proving a complete characterization of the positive and negative cases, we {\em necessarily} have to uncover all the important algorithmic ideas relevant to the family of problems we study. Indeed, our goal was not to prove a result specific to the kernelization of thin bipartite graphs, but it turned out that one cannot avoid proving this result in a complete characterization. The negative part of Theorem~\ref{theorem:intro:packing} shows that these two algorithmic ingredients (handling small components and thin bipartite graphs) cover all the relevant algorithmic ideas and {\em any} hereditary class $\F$ that cannot be handled by these ideas leads to a hard problem. For \FSubgraphTest, we first prove a dichotomy theorem characterizing the randomized polyno\-mial-time solvable and NP-hard cases. We say that $\F$ is {\em matching-splittable} if there is a constant $c$ such that every $H\in \F$ has a set $S$ of at most $c$ vertices such that every component of $H- S$ has at most 2 vertices. \begin{restatable}{\reabctheorem}{restateintropolysubgraph}\label{theorem:intro:polysubgraph} Let $\F$ be a hereditary class of graphs. If $\F$ is matching-splittable, then \FSubgraphTest can be solved in randomized polynomial time. If $\F$ does not have this property, then \FSubgraphTest is NP-hard. \end{restatable} The reason why randomization appears in Theorem~\ref{theorem:intro:polysubgraph} is the following. Given graphs $G$ and $H\in \F$, first we try every possible location where the set $S\subseteq V(H)$ can appear in $V(G)$ in a solution; as $|S|\le c$, there are $|V(G)|^c$ possibilities to try. Having fixed the location of $S$, we need to locate every component of $H-S$. As each such component is an edge or a single vertex, this looks like a matching problem, but here we have an additional restriction on how the endpoints of the edges should be attached to $S$. We can encode these neighborhood conditions using a bounded number of colors and get essentially a colored matching problem, which can be solved in randomized polynomial time using the algorithm of Mulmuley, Vazirani, and Vazirani \cite{MulmuleyVV87} for finding perfect matchings of exactly a certain weight. The negative side of Theorem~\ref{theorem:intro:polysubgraph} can be obtained by observing (using an application of Ramsey arguments) that if $\F$ is not matching-splittable, then $\F$ contains all cliques, all bicliques, all disjoint unions of triangles, or all disjoint unions of length-two paths; in each case, the problem is NP-hard. The authors are somewhat puzzled that the clean characterization of Theorem~\ref{theorem:intro:polysubgraph} has apparently not been observed so far in the literature: it is about the classical question of polynomial-time solvability of finding subgraphs and the proof uses techniques that are decades old. We may attribute this to the fact that while dichotomy theorems for fixed classes \F of graphs exist (e.g., \cite{DBLP:conf/icalp/ChenTW08,DBLP:journals/tcs/DalmauJ04,1206036,380867,DBLP:journals/tcs/KhotR02,LewisY80,DBLP:journals/siamcomp/Yannakakis81a}), perhaps it is not yet widely realized that such results are possible and aiming for them is a doable goal. We hope our paper contributes to the more widespread recognition of the feasibility of this line of research. In Theorem~\ref{theorem:intro:packing}, we have observed that Turing kernels are not more powerful than many-one kernels for \kFPacking. The situation is different for \kFSubgraphTest: there are classes $\F$ for which \kFSubgraphTest admits a polynomial Turing kernel, but has no polynomial many-one kernel, unless \containment. We characterize the classes $\F$ that admit polynomial Turing kernels the following way. We say that a graph $H$ is $(a,b,c,d)$-splittable, if there is a set $S$ of at most $c$ vertices such that every component of $H-S$ either has size at most $a$ or is a~$b$-thin bipartite graph with the additional restriction that the closed neighborhoods of all but $d$ vertices are universal to~$S$ (see Section~\ref{sec:char-prop} for details). \begin{restatable}{\reabctheorem}{restatemainsubgraph} \label{theorem:intro:turingsubgraph} Let $\F$ be a hereditary class of graphs. If there are $a,b,c,d\ge 0$ such that every $H\in \F$ is $(a,b,c,d)$-splittable, then \kFSubgraphTest admits a polynomial Turing kernel. If $\F$ does not have this property, then \kFSubgraphTest is \textup{WK[1]}-hard, \textup{W[1]}-hard, or \kPath-hard. \end{restatable} In the algorithmic part of Theorem~\ref{theorem:intro:turingsubgraph}, the first step is to guess the location of the set $S\subseteq V(H)$ in $V(G)$, giving $|V(G)|^c$ possibilities (this is the reason why in general our Turing kernel is not a many-one kernel). For each guess, locating the components of $H- S$ in $G$ is similar to Theorem~\ref{theorem:intro:packing}, as we have to handle small components and thin bipartite components, but here we have the additional technicality that we have to ensure that these components are attached to $S$ in a certain way. For many-one kernels, we do not have a characterization similar to Theorem~\ref{theorem:intro:turingsubgraph}. We present some concrete positive and negative results showing that a complete characterization of \kFSubgraphTest with respect to many-one kernels would be much more delicate than Theorem~\ref{theorem:intro:turingsubgraph}. The simple algorithmic idea used in Theorem~\ref{theorem:intro:turingsubgraph}, guessing the location of $S$, fails for many-one kernels and it seems that we have to make extreme efforts (whenever it is possible at all) to replace this step with adhoc arguments. \textbf{Our techniques.} The proofs of Theorems~\ref{theorem:intro:packing}--\ref{theorem:intro:turingsubgraph} all follow the same pattern. First, we define a certain graph-theoretic property and devise an algorithm for the case when $\F$ has this property. As described above, the algorithmic part of Theorem~\ref{theorem:intro:polysubgraph} is based on the randomized matching algorithm of Mulmuley, Vazirani, and Vazirani \cite{MulmuleyVV87}. For Theorems~\ref{theorem:intro:packing} and \ref{theorem:intro:turingsubgraph}, the algorithm is a marking procedure: for each component, we mark a bounded number of vertices such that we can always find a copy of this component using only these vertices even if the other components already occupy an unknown but small set of vertices. Therefore, if there is a solution, then there is a solution using only this set of marked vertices. The kernel is obtained by restricting the graph to this set of vertices. For small components, we use the Sunflower Lemma of Erd\H os and Rado \cite{MR22:2554} (similarly as it is used in the kernelization of other packing problems, cf.~\cite{DellM12}). For thin bipartite graphs, the marking procedure is a branching algorithm specifically designed for this class of graphs. At some point in the algorithm, we crucially use that the component is $b$-thin: we find a biclique with $b$ vertices on one side and many vertices on the other side, and then we argue that the component is a subgraph of this biclique. For the hardness results of Theorems~\ref{theorem:intro:packing}--\ref{theorem:intro:turingsubgraph}, first we prove that if $\F$ does not have the stated property, then $\F$ contains every graph from one of the basic families of hard graphs. These hard families include cliques, bicliques, paths, odd cycles with a high-degree vertex, and subdivided stars (see Section~\ref{sec:hard-families}). To prove that a hard family appears in $\F$, we use Ramsey results (including a recent path vs.~induced path vs.~biclique result of Atminas, Lozin, and Razgon~\cite{AtminasLR12}) and a graph-theoretic analysis of what, for example, a large nonbipartite graph without large cliques and long induced paths can look like. For each hard family, we then claim a lower bound on the problem. Most of these lower bounds take the form of a relatively standard polynomial-parameter transformation from \textsc{Set Cover} parameterized by the size of the universe; here the value of our contribution is not in the details of the reduction, but in realizing that these are the hard families of graphs whose hardness exhaustively explain the hard cases of the problem. The basic technique to obtain negative evidence for the existence of many-one kernels is the method of \emph{OR-cross-composition}~\cite{BodlaenderJK14}, which refines the original OR-composition framework~\cite{BodlaenderDFH09}. An OR-cross-composition of a classical problem~$L$ into a parameterized problem~$\Q$ is a polynomial-time embedding of a series of~$t$ length-$n$ instances~$x_1, \ldots, x_t$ of~$L$ into a single instance~$x^*$ of~$\Q$ with parameter value poly($n$), such that~$x^* \in \Q \Leftrightarrow \bigvee _{i=1}^t x_i \in L$. If~$L$ is NP-hard, such a construction is known to rule out the existence of polynomial kernels for \Q under the assumption that \ncontainment. The negative results that we present for the existence of many-one kernels for \FSubgraphTest use a specific form of this technique that we name {\em OR-cross-composition by reduction with a canonical template.} The idea is to start from an NP-hard graph problem~$L$ for which a family of polynomial-size \emph{canonical template graphs} exists, such that for every~$n$, the instances of length~$n$ are induced subgraphs of the $n$-th graph in this family. This allows length-$n$ inputs~$x_1, \ldots, x_t$ to be merged into one through their common canonical supergraph of size poly($n$), as opposed to the trivial~$t \cdot n$, which facilitates an OR-cross-composition. Canonical template graphs were first used for this purpose by Bodlaender et al.~\cite[Theorem 11]{BodlaenderJK12c} to prove a kernel lower bound for a structural parameterization of \textsc{Path with Forbidden Pairs}. \section{Outline}\label{sec:outline} In this section we present a more detailed overview of the results of the paper. We also describe the main technical parts of the proofs. The proofs of Theorems~\ref{theorem:intro:packing}--\ref{theorem:intro:turingsubgraph} all follow the same pattern: \begin{enumerate*} \item We define the property separating the positive and negative cases. \item We prove an algorithmic result for the positive cases. \item We prove a purely combinatorial result stating that if a class $\F$ does not satisfy the property, then $\F$ is a superset of one of the classes appearing on a short list of basic hard classes. \item We prove a hardness result for each basic hard class on the list. \end{enumerate*} The structure of this section follows these steps: for each step, we go through the relevant definitions and state the results proved later in the paper. \subsection{Characterizing properties} \label{sec:char-prop}We say that a graph is {\em $c$-matching-splittable} if there is a set $S\subseteq V(H)$ of at most $c$ vertices such that every component of $H- S$ has at most two vertices. We say that a class $\F$ of graphs is $c$-matching-splittable if every $H\in \F$ has this property, and we say that $\F$ is matching-splittable if $\F$ is $c$-matching-splittable for some $c\ge 0$. In Theorem~\ref{theorem:intro:polysubgraph}, this is the condition for randomized polynomial-time solvability. Clearly, a matching is 0-matching-splittable and a matching plus a universal vertex is 1-matching-splittable. On the other hand, the class containing the disjoint unions of arbitrarily many triangles is not $c$-matching-splittable for any $c\ge 0$, as $S$ would need to contain at least one vertex from each triangle. In Theorem~\ref{theorem:intro:packing}, the condition that we need is that every component is either small or a thin bipartite graph. We say that a graph $H$ is {\em $a$-small/$b$-thin} if every component of $H$ has at most~$a$ vertices or is a $b$-thin bipartite graph (that is, a bipartite graph with one of the partite classes having size at most $b$). Note that $H$ can have both types of components. For example, if $H$ is the disjoint union of an arbitrary number of triangles and stars of arbitrary size, then it is 3-small/1-thin. We say that class $\F$ is $a$-small/$b$-thin if every graph $H\in \F$ has this property and say that $\F$ is small/thin if it is $a$-small/$b$-thin for some $a,b\ge 0$. The characterization property that we need for Theorem~\ref{theorem:intro:turingsubgraph} is a somewhat technical generalization of being $a$-small/$b$-thin. \begin{definition}\label{def:splittable} We say that a graph $H$ is \emph{$(a,b,c,d)$-splittable} if it has a vertex set~$S \subseteq V(H)$ of size at most~$c$ such that: \begin{enumerate} \item each connected component of~$H - S$ on more than~$a$ vertices is bipartite and has a partite class of size at most~$b$, and \item in each connected component~$C$ of~$H - S$, the number of vertices whose closed neighborhood in~$G[C]$ is \emph{not} universal to~$N_H(C) \cap S$ is at most~$d$. \end{enumerate} We say that such a set~$S \subseteq V(H)$ \emph{realizes} the~$(a,b,c,d)$-split of~$H$. Family~\F is $(a,b,c,d)$-splittable if every $H\in \F$ is a $(a,b,c,d)$-splittable. Family~\F is splittable if there are constants~$a,b,c,d$ such that~\F is $(a,b,c,d)$-splittable. \end{definition} \noindent Observe that being $a$-small/$b$-thin is exactly the same as being $(a,b,0,0)$-splittable and being $c$-matching-splittable is exactly the same as being $(2,0,c,2)$-splittable. We prefer to use the terms $a$-small/$b$-thin and $c$-matching-splittable for these special cases, as they are more descriptive. Given an $a$-small/$b$-thin graph $H$, adding a set $S$ of $c$ universal vertices results in an $(a,b,c,0)$-splittable graph $H'$. If $C$ is a component of $H$ having at most $a$ vertices and we remove from $H'$ any set of edges between $C$ and $S$, then the resulting graph is $(a,b,c,a)$-splittable. The closed neighborhoods of the $a$ vertices in~$C$ may no longer be universal to~$N_H(C) \cap S$ after the edge removals, which is compensated by the fourth entry in the tuple. Let now $C$ be a $b$-thin bipartite component of $H$, let $A$ be the smaller side and let $B$ be the larger side of $C$. Observe that Definition~\ref{def:splittable} not only requires that all but $d$ vertices of $C$ are universal to $N_H(C)\cap S$, but even the closed neighborhoods in~$G[C]$ have to be universal. Therefore, removing even a single edge between a vertex $v$ of $C$ and $S$ can ruin the property, as it ``contaminates'' all the neighbors of $v$. If we remove a single edge between some $x\in B$ and $S$, then the graph is still $(a,b,c,b+1)$-splittable: there are at most $b+1$ vertices in $C$ whose neighborhood is not universal to $S$, namely $x$ and some of the vertices of $A$. On the other hand, if we remove a single edge between some $y\in A$ and $S$, then the graph may not be $(a,b,c,d)$-splittable for arbitrary large $d$: if $y$ has degree $d$, then $y$ and all its neighbors have the property that their closed neighborhoods are not universal to $S$. Note that the definition {\em does not} require that the closed neighborhood of (all but $d$ of) the vertices are universal to $S$, it requires universality only to $N_H(C)\cap S$. Suppose that $H_1$ and $H_2$ are two graphs with $S_i$ realizing an $(a,b,c,d)$-split of $H_i$ for $i=1,2$. The disjoint union of $H_1$ and $H_2$ is $(a,b,2c,d)$-splittable, as realized by $S_1\cup S_2$: the vertices in a $b$-thin component of $H_1$ need to be universal only to (a certain part of) $S_1$, as $C$ has no edge to $S_2$. \subsection{Algorithms} In the algorithmic part of Theorem~\ref{theorem:intro:polysubgraph}, we need to solve \SubgraphTest in the case that~$H$ is $c$-matching-splittable for some set $S$ of at most $c$ vertices. As described in the introduction, we guess the location of $S$ and then solve the resulting constrained matching problem. The main technical engine in the algorithm is the classic algebraic matching algorithm due to Mulmuley, Vazirani, and Vazirani~\cite{MulmuleyVV87}. It can be used to obtain randomized algorithms for various colored versions of matching (see, for example, \cite{DBLP:journals/ipl/Marx04,MarxP14}). We need the following variant. \begin{restatable}{\retheorem}{restatecoloredmatching} \label{theorem:coloredmatching} Given a multigraph $G$ with a (not necessary proper) coloring of the edges with a set $C$ of colors and function $f:C\to \mathbb{Z}^+$, there is a randomized algorithm with false negatives that decides in time $(|V(G)|+|E(G)|)^{O(|C|)}$ if $G$ has a matching containing exactly $f(i)$ edges of color $i$ for every $i\in C$. \end{restatable} By a randomized algorithm with false negatives, we mean an algorithm that is always correct on \no-instances, but which may incorrectly reject a \yes-instance with probability at most~$\frac{1}{2}$. Equipped with Theorem~\ref{theorem:coloredmatching}, we prove the algorithmic part of Theorem~\ref{theorem:intro:polysubgraph} in Section \ref{section:polynomialvsnpcomplete:upperbounds}. \begin{restatable}{\retheorem}{restatesubgraphtestalg} \label{theorem:fsubgraphtest:alg} \FSubgraphTest is (randomized) polynomial-time solvable if $\F$ is matching-splittable. \end{restatable} The polynomial kernel in the positive part of Theorem~\ref{theorem:intro:packing} is obtained by a marking procedure that finds a polynomially bounded subset of vertices in $G$ that surely contains a solution, if a solution exists at all. Let us first explain briefly how the standard technique of sunflowers can be used for this marking procedure if every component of $H$ has at most $a$ vertices. We need the Sunflower Lemma of Erd\H os and Rado \cite{MR22:2554}. A collection $\S$ of sets is called a {\em sunflower} if the pairwise intersection $S_1\cap S_2$ is the same set $C$ for any two distinct $S_1,S_2\in \S$. Then this intersection $C$ is the {\em core} of the sunflower; the sets $S\setminus C$ for $S\in \S$ are the {\em petals} of the sunflower. \begin{lemma}[{\cite{MR22:2554}, cf.~\cite[Lemma 9.7]{FlumG06}}] \label{lemma:sunflowers} Let $k$ and $m$ be nonnegative integers and let~$\S$ be a system of sets of size at most~$m$ over a universe~$U$. If~$|\S| \geq m!(k-1)^{m}$, then there is a sunflower in~$\S$ with~$k$ petals. Furthermore, for every fixed~$m$ there is an algorithm that computes such a sunflower in time polynomial in~$(k + |\S|)$. \end{lemma} Let $H$, $G$, and $t\ge 1$ form an instance of \Packing; the solution we are looking for has $k:=t\cdot |V(H)|$ vertices. Let $C$ be a component of $H$ having size at most $a$. First, we enumerate every subset of $|V(C)|\le a$ vertices in $G$ where $C$ appears; the length of this list is polynomial in the size of $G$ if $a$ is a fixed constant. We would like to reduce the length of this list: we would like to have a shorter list of candidate locations where $C$ can appear in a solution, such that the length of the list is polynomially bounded in $k$. We argue the following way. As long as the length of the list is at least $a!(k+1)^a$, we can find a sunflower with $k+2$ petals among the sets in the list. We claim that we can choose any set $S$ from this sunflower and throw it out of the list. Suppose that there is a solution where the component $C$ is mapped exactly to this set $S\subseteq V(G)$. As the solution uses only $k$ vertices of $G$ and the petals of the sunflower are disjoint, there is another set $S'$ among the remaining $k+1$ sets of the sunflower whose petal is disjoint from the solution. Therefore, we can modify the solution such that $C$ is mapped to $S'$ instead of $S$, which means that the set $S$ cannot be essential to the solution and can be safely removed from the list of candidate locations for $C$. Repeating this argument, we eventually get a list of at most $a!(k+1)^a$ candidate locations for each component of $H$, thus we can reduce the problem to an induced subgraph of $G$ whose size, for a fixed constant $a$, is polynomial in $k$. If $H$ has $b$-thin components, then the Sunflower Lemma cannot be applied, as the size of such a component can be arbitrarily large (and it is the size of the component that appears in the exponent in the argument above). Therefore, in Section~\ref{sec:repr-sets-thin}, we develop a marking procedure specifically designed for thin bipartite graphs. As an illustration, we present here the main idea on the special case of packing thin bicliques, that is, on graphs $K_{b,\ell}$ for some fixed $b\ge 1$. The crucial ingredient for the kernel for biclique packing is the following lemma. \begin{lemma} \label{lemma:bicliquemarking:specialcase} For every fixed~$b$ there is a polynomial-time algorithm that, given a graph~$G$ and integers~$\ell > b$ and~$k \geq \ell + b$, computes a set~$X$ of size~$\Oh(k^{4b})$ such that for every~$Z \subseteq V(G)$ of size at most~$k$, if~$G - Z$ contains a $K_{b,\ell}$ subgraph, then~$G[X] - Z$ contains a $K_{b,\ell}$ subgraph. \end{lemma} Before proving the lemma, we show how it leads to a polynomial kernel for biclique packing. To reduce the size of an instance that asks whether~$G$ contains~$t$ disjoint $K_{b,\ell}$ subgraphs for~$\ell > b$, we define~$k := t \cdot (b + \ell)$ and invoke the lemma to compute a set~$X$ of size~$\Oh(k^{4b})$. We then output~$G[X]$ as the kernelized instance. If~$G$ contains a packing of~$t$ disjoint $K_{b,\ell}$ subgraphs, then while the packing contains a biclique~$C$ using a vertex in~$V(G) \setminus X$, we let~$Z$ be the~$(t-1)(b + \ell)$ other vertices in the packing, apply the guarantee of the lemma to find a biclique model~$C'$ in~$G[X]$ avoiding~$Z$, and replace~$C$ in the packing by~$C'$. Iterating the argument results in a packing of bicliques in~$G[X]$, proving that the reduced instance is equivalent to the original one. To facilitate a recursive algorithm, we actually prove a generalization of Lemma~\ref{lemma:bicliquemarking:specialcase}. To state the generalization we need the following terminology. For disjoint sets~$A', B' \subseteq V(G)$ and~$\ell > b$ we say that a~$K_{b,\ell}$ subgraph in~$G$ extends~$(A',B')$ if the side-$b$ partite class is a superset of~$A'$ and the size-$\ell$ partite class is a superset of~$B'$. \begin{lemma} \label{lemma:bicliquemarking:specialcase:general} For every fixed~$b$ there is a polynomial-time algorithm that, given a graph~$G$, integers~$\ell > b$ and~$k \geq \ell + b$, and disjoint sets~$A', B' \subseteq V(G)$ of size at most~$b$, computes a set~$X$ of size at most~$(3k^2)^{2b - |A' \cup B'|}$ such that for every~$Z \subseteq V(G)$ of size at most~$k$, if~$G - Z$ contains a $K_{b,\ell}$ subgraph that extends~$(A',B')$, then~$G[X] - Z$ contains a $K_{b,\ell}$ subgraph. \end{lemma} \begin{proof} The main idea behind the algorithm is to make progress in recursive calls by increasing the size of~$A' \cup B'$, thereby restricting the type of bicliques that have to be preserved in the set~$X$. Throughout the proof we use the fact that if~$Z \subseteq V(G)$ and there is a $K_{b,\ell}$-subgraph in~$G - Z$ that extends~$(A',B')$, then~$Z \cap (A' \cup B') = \emptyset$, the size-$b$ partite class consists of common neighbors of~$B'$, while the size-$\ell$ partite class consists of common neighbors of~$A'$. Let us point out that the lemma requires that $G[X]-Z$ contains an $K_{b,\ell}$-subgraph, but it does not require it to extend $(A',B')$. \textbf{Case 1.} If~$|A'| = b$, then we choose~$X$ as~$A' \cup B'$ together with~$k + \ell$ common neighbors of~$A'$ (or less, if there are fewer), for a total size of at most~$2b + (k + \ell) \leq 3k$. Let~$Z \subseteq V(G)$ have size at most~$k$. If there is a $K_{b,\ell}$-subgraph~$H$ in~$G - Z$ that extends~$(A',B')$, then all vertices in the size-$\ell$ partite class are common neighbors of~$A'$. If all vertices of~$H$ are contained in~$G[X]$, then the biclique subgraph~$H$ also exists in~$G[X]-Z$. If not, then the set~$A'$ had at least~$k + \ell$ common neighbors (otherwise they were all preserved in~$X$). Since~$Z$ contains at most~$k$ of them, any~$\ell$ of the remaining vertices in~$X$ combines with~$A'$ to form a~$K_{b,\ell}$-subgraph in~$G[X] - Z$. \textbf{Case 2.a.} If~$|B'| = b$,~$|A'| < b$, and the set~$B'$ has at least~$k + \ell$ common neighbors, then we choose~$X$ containing~$k+\ell$ of these common neighbors together with~$B'$ itself. For any~$Z \subseteq V(G)$ of size at most~$k$, if a biclique extending~$(A',B')$ exists in~$G - Z$ then~$Z$ avoids at least~$\ell$ common neighbors of~$B'$ in~$X$. Together with~$B'$, these form a~$K_{b,\ell}$ subgraph in~$G[X] - Z$. Note that this $K_{b,\ell}$ does not extend $(A',B')$, but this is not required by the lemma. \textbf{Case 2.b.} If~$|B'| = b$,~$|A'| < b$, and the set~$B'$ has less than~$k + \ell \leq 2k$ common neighbors~$T := \bigcap _{v \in B'} N_G(v)$, then a $K_{b,\ell}$-subgraph extending~$(A',B')$ has its size-$b$ side within~$T$. For each~$a \in T \setminus (A' \cup B')$, add~$a$ to~$A'$ and recurse. Let~$X$ be the union of the recursively computed sets. If there is a biclique in~$G - Z$ extending~$(A',B')$, then there is an~$a \in T \setminus (A' \cup B')$ such that it extends~$(A' \cup \{a\}, B')$, and the correctness guarantee for that recursive call yields a biclique in~$G[X] - Z$. The measure~$2b - |A' \cup B'|$ drops in each recursive call and we recurse on at most~$2k$ instances, giving a bound of~$2k \cdot (3k^2)^{2b - |A' \cup B'| - 1} \leq (3k^2)^{2b - |A' \cup B'|}$ on~$|X|$. \textbf{Case 3.} In the remaining cases we have~$|A'|, |B'| < b$. We greedily compute a maximal set of $K_{b,\ell}$ subgraphs that extend~$(A',B')$ and pairwise intersect only in~$A' \cup B'$. Since~$b$ is constant, this can be done in polynomial time by guessing all possible locations for the remaining vertices in the size-$b$ partite class and testing whether the resulting vertices are adjacent to~$B'$ and have sufficient common neighbors to realize the other partite class. Two things can happen. \textbf{Case 3.a.} If we find~$k + 1$ distinct $K_{b,\ell}$ subgraphs that pairwise intersect only in~$(A',B')$, then we output~$X$ containing the union of these subgraphs, which has size at most~$(k+1)(\ell + b) \leq 2k^2$. If a $K_{b,\ell}$-subgraph extending~$(A',B')$ exists in~$G - Z$ for some~$Z \subseteq V(G)$ of size~$k$, then~$Z$ intersects at most~$k$ of the extensions. Hence one extension avoids~$Z$ and combines with~$A',B'$ to form a~$K_{b,\ell}$-subgraph in~$G[X] - Z$. \textbf{Case 3.b.} If there are at most~$k$ of such extensions, then let~$T$ contain the at most~$k (\ell + b) \leq k^2$ vertices in their union. By the maximality of the packing, any extension of~$(A',B')$ uses a vertex in~$T \setminus (A' \cup B')$. For each~$v \in T \setminus (A' \cup B')$, recurse twice: once for adding~$v$ to~$A'$ and once for adding~$v$ to~$B'$. We let~$X$ be the union of the recursively computed sets. If there is a $K_{b,\ell}$ subgraph in~$G - Z$ for some~$Z \subseteq V(G)$ of size at most~$k$, then it extends~$(A' \cup \{v\}, B')$ or~$(A', B' \cup \{v\})$ for some~$v \in T \setminus (A' \cup B')$. The correctness guarantee for that branch of the recursion guarantees the existence of~$K_{b,\ell}$ in~$G[X] - Z$. As the measure~$2b - |A' \cup B'|$ drops in each recursive call, while we branch in at most~$2|T| \leq 2k(\ell + b) \leq 2k^2$ directions, the size of~$X$ is bounded by~$2k^2 \cdot (3k^2)^{2b - |A' \cup B'| - 1} \leq (3k^2)^{2b - |A' \cup B'|}$. \end{proof} The generalization from $b$-thin bicliques to general $b$-thin bipartite graphs makes the scheme described above much more technical. Let us point out that the large side of a $b$-thin bipartite graph can be partitioned into at most $2^b$ classes according to its neighborhood in the small side. Therefore, intuitively, a $b$-thin bipartite graph can be seen as $2^b$ different $b$-thin bicliques joined together, which makes it plausible that such a generalization exists. \begin{restatable}{\retheorem}{restatepackingkernel}\label{theorem:kernel:packing} If $\F$ is a hereditary class of graphs that is small/thin, then \FPacking admits a polynomial many-one kernel. \end{restatable} For the algorithmic part of Theorem~\ref{theorem:intro:turingsubgraph}, we have to guess the location of the set $S$ realizing the $(a,b,c,d)$-split and then take into account the universality restrictions. This introduces another layer of technical difficulties, but no new conceptual ideas are needed. Moreover, because of this guessing step, the kernel is no longer many-one, but it is a Turing kernel. \begin{restatable}{\retheorem}{restatesubgraphturing} \label{theorem:kernel:subgraph} If $\F$ is a hereditary class of graphs that is splittable, then \FSubgraphTest admits a polynomial Turing kernel. \end{restatable} \subsection{Hard families} \label{sec:hard-families} We define several specific classes of graphs and show hardness results for these classes. Then we show that if a class does not have the property of, say, being splittable, then it is a superset of at least one hard class, hence hardness follows for every class that does not have this property. First, we define the following graphs (see Figure~\ref{fig:basic} on page~\pageref{fig:basic}). \begin{itemize}[noitemsep] \item $\Gpath{\ell}$ is the path of length $\ell$, which consists of~$\ell$ edges and ~$\ell + 1$ vertices. It is sometimes denoted~$P_{\ell+1}$ for brevity. \item $\Gclique{n}$ is the clique on~$n$ vertices (while describing hard families, we use $\Gclique{n}$ instead of the more standard $K_n$ for consistency of notation). \item $\Gbiclique{n}$ is the balanced biclique~$K_{n,n}$ on $n+n$ vertices. \item $\Gdoublebroom{s}{n}$ is obtained from a length-$s$ path by adding $n$ pendant vertices to each of the two endpoints of the path. \item $\Goperahouse{s}{n}$ is obtained from a length-$s$ path by adding $n$ vertices that are adjacent to both endpoints of the path. \item $\Gfountain{s}{n}$ is obtained from a length-$s$ cycle by adding $n$ pendant vertices to one vertex on the cycle. \item $\Glongfountain{s}{t}{n}$ is obtained from a length-$s$ cycle by adding a path of length~$t$, identifying one endpoint with a vertex on the cycle and adding $n$ pendant vertices to the other endpoint. \item $\Gsubdivstar{n}$ is obtained from a star with $n$ leaves by subdividing each edge once. \item $\Gsubdivtree{s}{n}$ is obtained from a star with $n$ leaves by subdividing each edge $s-1$ times and attaching $n$ pendant vertices to each leaf. \item $\Gdiamondfan{n}$ is obtained from $n$ copies of $K_{2,n}$ by taking one degree-$n$ vertex from each copy and identifying them into a single vertex. \end{itemize} \begin{figure}[t] \begin{center} {\small \svg{\linewidth}{families}} \end{center} \caption{Basic families of graphs.}\label{fig:basic} \end{figure} We can define families of these graphs the obvious way: \[ \begin{array}{ll} \Fpath=\{\Gpath{i}\mid i\ge 1\}& \Fclique=\{\Gclique{i}\mid i\ge 1\}\\ \Fbiclique=\{\Gbiclique{i}\mid i\ge 1\}& \Fdoublebroom{s}=\{\Gdoublebroom{s}{i}\mid i\ge 1\}\\ \Ffountain{s}=\{\Gfountain{s}{i}\mid i\ge 1\}& \Flongfountain{s}{t}=\{\Glongfountain{s}{t}{i}\mid i\ge 1\}\\ \Foperahouse{s}=\{\Goperahouse{s}{i}\mid i\ge 1\}& \Fsubdivstar=\{\Gsubdivstar{i}\mid i\ge 1\}\\ \Fsubdivtree{s}=\{\Gsubdivtree{s}{i}\mid i\ge 1\}& \Fdiamondfan=\{\Gdiamondfan{i}\mid i\ge 1\} \end{array} \] To prove that a hard family is contained in every class not satisfying a certain property, we use arguments based on Ramsey theory. The following lemma (proved in Section~\ref{sec:polyn-time-solv}) characterizes hereditary classes that are not matching-splittable. We define $n\cdot H$ to be the graph that contains $n$ disjoint copies of $H$. (Recall that $P_3$ is the path on 3 vertices.) \begin{restatable}{\retheorem}{restateramseymatching} \label{theorem:ramsey:matchingsplittable} Let $\F$ be a hereditary graph family that is not matching splittable. Then at least one of the following holds: \begin{enumerate*} \item $\F$ is a superset of \Fclique. \item $\F$ is a superset of \Fbiclique. \item $\F$ contains $n\cdot K_3$ for every $n\ge 1$. \item $\F$ contains $n\cdot P_3$ for every $n\ge 1$. \end{enumerate*} \end{restatable} Observe that Theorem~\ref{theorem:ramsey:matchingsplittable} is a tight characterization of matching-splittable graphs: the converse statement is also true, that is, if any of the four statements is true for $\F$, then it is not matching-splittable. Clearly, large cliques and large bicliques are not $c$-matching-splittable for constant $c$. Moreover, if every component of a graph has three vertices (that is, it is either a $K_3$ or $P_3$), then at least one vertex has to be deleted from each component to decrease the size of every component to at most two vertices, hence $\F$ cannot be $c$-matching-splittable for constant $c$ in the last two cases either. In Section~\ref{sec:packing-ramsey}, we characterize hereditary classes that are not small/thin. \begin{restatable}{\retheorem}{restateramseysmallthin} \label{theorem:ramsey:1} Let~$\F$ be a hereditary graph family that is not small/thin. Then $\F$ is a superset of at least one of the following families: \begin{enumerate*} \item \Fpath, \item \Fclique, \item \Fbiclique, \item \Ffountain{s} for some odd integer $s\ge 3$, \item \Flongfountain{s}{t} for some odd integer $s\ge 3$ and integer $t\ge 1$, \item \Foperahouse{s} for some odd integer $s\ge 1$, \item \Fsubdivstar, or \item \Fdoublebroom{s} for some odd integer $s\ge 1$. \end{enumerate*} \end{restatable} Again, the characterization is tight: we can observe that if $\F$ is a superset of any of these families, then there is no $a,b\ge 0$ such that $\F$ is $a$-small/$b$-thin. Note that we cannot leave out any of the eight items from the list: the hereditary closure of, say, \Flongfountain{5}{2} is not the superset of any of the classes described in the remaining seven items. Finally, in Section~\ref{sec:subgraph-ramsey}, we characterize graphs that are not splittable. \begin{restatable}{\retheorem}{restateramseysplittable} \label{theorem:ramsey:separator} Let~$\F$ be a hereditary graph family that is not splittable. Then at least one of the following holds: \begin{enumerate*} \item $\F$ is a superset of $\Fpath$, \item $\F$ is a superset of $\Fclique$, \item $\F$ is a superset of $\Fbiclique$, \item $\F$ contains $n\cdot \Gsubdivstar{n}$ for every $n\ge 1$, \item there is an odd $s\ge 3$ such that $\F$ contains $n\cdot \Gfountain{s}{n}$ for every $n\ge 1$, \item there is an odd $s\ge 1$ such that $\F$ contains $n\cdot \Goperahouse{s}{n}$ for every $n\ge 1$, \item there is an odd $s\ge 1$ such that $\F$ contains $n\cdot \Gdoublebroom{s}{n}$ for every $n\ge 1$, \item there is an odd $s\ge 3$ and arbitrary $t\ge 1$ such that $\F$ contains $n\cdot \Glongfountain{s}{t}{n}$ for every $n\ge 1$, \item $\F$ is a superset of $\Fsubdivtree{s}$ for some integer $s\ge 1$, or \item $\F$ is a superset of $\Fdiamondfan$. \end{enumerate*} \end{restatable} We can again verify that the characterization is tight. In particular, let us show that $\Gsubdivtree{s}{n}$ is not $(a,b,c,d)$-splittable if $n>a+b+c+d$. Suppose that $S$ realizes the $(a,b,c,d)$-split. As the graph is not $b$-thin and has more than $a$ vertices, we have that $S$ is not empty. By the pigeonhole principle, there is a vertex $v$ with degree $n+1$ that is not in $S$, and none of its degree-1 neighbors are in $S$ either. Then $v$ is in a component of size at least $n+1>a$ that contains at least $n>d$ vertices that have no neighbors in $S$. Similarly, suppose that $S$ realizes an $(a,b,c,d)$-split of $\Gdiamondfan{n}$ for $n>a+b+c+d$. Again, $S$ is not empty. By the pigeonhole principle, there is a degree-$n$ vertex $v$ that is not in $S$ and has no neighbor in $S$. The component of this vertex has size more than $a$ and the component has more than $d$ vertices (namely, every neighbor of $v$) whose closed neighborhood is not universal to $S$. \subsection{Hardness proofs} \label{section:outline:hardness} Let us review the concrete hardness results that we prove, which, by the combinatorial characterizations in Theorems \ref{theorem:ramsey:matchingsplittable}--\ref{theorem:ramsey:separator}, prove the negative parts of Theorems~\ref{theorem:intro:packing}--\ref{theorem:intro:turingsubgraph}. As mentioned above, Kirkpatrick and Hell \cite{KirkpatrickH78} fully characterized the polynomial-time solvable cases of \HPacking. \begin{theorem}[\cite{KirkpatrickH78}] \label{theorem:fpacking:pvsnp} \HPacking is polynomial-time solvable if every connected component of $H$ has at most two vertices and NP-complete otherwise. \end{theorem} It follows from Theorem~\ref{theorem:fpacking:pvsnp} that \FSubgraphTest is NP-hard if $\F$ contains $n\cdot K_3$ for every $n\ge 1$ or if $\F$ contains $n\cdot P_3$ for every $n\ge 1$, as then the problem is more general than $K_3$-\Packing or $P_3$-\Packing, respectively. Also, \FSubgraphTest is NP-hard if $\F$ contains every clique~\cite[GT7]{GareyJ79} (it generalizes \textsc{Clique}) or if $\F$ contains every biclique~\cite[GT24]{GareyJ79}. For kernelization lower bounds, observe first that if~$\F$ contains every clique, then \FPacking and \FSubgraphTest are clearly W[1]-hard~\cite[Theorem 21.2.4]{DowneyF13} and therefore do not admit a (Turing) kernel of any size, unless FPT~$=$~W[1] and the Exponential Time Hypothesis fails~\cite[Chapter 29]{DowneyF13}. A recent result of Lin~\cite{Lin14} shows that if~$\F$ contains every biclique, then \FPacking and \FSubgraphTest are also W[1]-hard. Since the parameterized \textsc{Clique} and \textsc{Biclique} problems are NP-hard and OR-compositional~\cite{BodlaenderDFH09}, it follows from standard kernelization lower bound machinery that if \FPacking or \FSubgraphTest has a polynomial (many-one) kernel when~$\F$ contains every clique or biclique, then \containment. To complete the proof of the negative parts of Theorems~\ref{theorem:intro:packing} and \ref{theorem:intro:turingsubgraph}, we prove the following two sets of \textup{WK[1]}-hardness results. \begin{restatable}{\retheorem}{restatepackinglower} \label{theorem:packing:lowerbounds} The \FPacking problem is \textup{WK[1]}-hard under polynomial-parameter transformations if $\F$ is a superset of any of the following families: \begin{enumerate*} \item $\Fsubdivstar$, \item $\Flongfountain{s}{t}$ for some integer~$t \geq 1$ and some \emph{odd} integer~$s \geq 3$, \item $\Fdoublebroom{s}$ for some odd integer $s \geq 1$, \item $\Ffountain{s}$ for some odd integer $s \geq 3$, or \item $\Foperahouse{s}$ for some odd integer $s \geq 1$. \end{enumerate*} \end{restatable} \begin{restatable}{\retheorem}{restatesubgraphlower}\label{theorem:subgraphtest:lowerbounds} The \FSubgraphTest problem is \textup{WK[1]}-hard under polynomial-parameter transformations if $\F$ is a superset of any of the following families: \begin{enumerate*} \item $\Fdiamondfan$, or \item $\Fsubdivtree{s}$ for some integer $s$. \end{enumerate*} \end{restatable} All \textup{WK[1]}-hardness proofs are by reduction from \nRegularExactSetCover, where the parameter equals the size of the universe on which the set system is defined. The uniform variant, in which all sets have the same size, is particularly useful for proving these results. We prove the \textup{WK[1]}-hardness of the problem by a two-stage transformation from \nExactSetCover~\cite{HermelinKSWW13}, first introducing a small number of new elements to ensure that solutions exist that contain a prescribed number of sets, and then using this knowledge to introduce another small number of elements that can be added to the sets to make the system uniform. \subsection{Many-one kernels} We do not have a complete characterization of the existence of many-one kernels for \FSubgraphTest. The authors believe that if such a characterization is possible, then it has to be significantly more delicate than the characterization of Turing kernels in Theorem~\ref{theorem:intro:turingsubgraph} and both the positive and the negative parts should involve a larger number of specific cases. We present two lower bounds and two upper bounds to show the difficulties that arise (see also Figure~\ref{fig:karp}). The following two theorems give the lower bounds. \begin{figure}[t] \begin{center} {\small \svg{\linewidth}{karp}} \caption{Illustrating the classes of graphs in Theorems~\ref{theorem:karp:lowerbound:onesubstar:manytriangles}--\ref{theorem:karp:subgraphkernel:starsandpaths}.} \label{fig:karp} \end{center} \end{figure} \begin{restatable}{\retheorem}{restatekarponesubstarmanytriangles} \label{theorem:karp:lowerbound:onesubstar:manytriangles} Let~\F be any hereditary graph family containing all graphs of the form~$H' + \ell \cdot K_3$, where~$\ell \geq 1$ and~$H' \in \Fsubdivstar$. Then \kFSubgraphTest does not admit a polynomial many-one kernel unless \containment. \end{restatable} \begin{restatable}{\retheorem}{restatekarptwostarmanyps} \label{theorem:karp:lowerbound:twosubstar:manyps} Let~\F be any hereditary graph family containing all graphs of the form~$H' + H'' + \ell \cdot P_3$, where~$\ell \geq 1$ and~$H', H'' \in \Fsubdivstar$. Then \kFSubgraphTest does not admit a polynomial many-one kernel unless \containment. \end{restatable} Observe that the graph families described by these theorems are~$(3,0,2,2)$-splittable: letting~$S$ contain the (at most two) centers of the subdivided stars, the connected components that remain after removing~$S$ have at most three vertices. Every leg of a subdivided star becomes a component of size two in which one of the vertices is universal to~$S$ and the other is not; hence the closed neighborhoods of the two vertices are not universal to~$S$. The \kFSubgraphTest problem for these families therefore has polynomial Turing kernels by Theorem~\ref{theorem:kernel:subgraph}, highlighting the difference between many-one and Turing kernelization for \kFSubgraphTest. The following two theorems give upper bounds. \begin{restatable}{\retheorem}{restatekarpfountaintriangles} \label{karp:subgraphkernel:fountainandtriangles} Let~\F be the hereditary closure of the family containing all graphs of the form~$H' + \ell \cdot K_3$, where~$\ell \geq 1$ and~$H' \in \Ffountain{3}$. Then \kFSubgraphTest admits a polynomial many-one kernel. \end{restatable} \begin{restatable}{\retheorem}{restatekarpstarspaths} \label{theorem:karp:subgraphkernel:starsandpaths} Let~\F be the hereditary closure of the family containing all graphs of the form~$H' + \ell \cdot P_3$, where~$\ell \geq 1$ and~$H' \in \Fsubdivstar$. Then \kFSubgraphTest admits a polynomial many-one kernel. \end{restatable} Comparing Theorem~\ref{theorem:karp:lowerbound:onesubstar:manytriangles} to Theorem~\ref{karp:subgraphkernel:fountainandtriangles}, we find that changing the type of the single large component from a subdivided star to a fountain crosses the threshold for the existence of a polynomial kernel, even though both types of graphs can be reduced to constant-size components by a single vertex deletion. Comparing Theorem~\ref{theorem:karp:lowerbound:twosubstar:manyps} to Theorem~\ref{theorem:karp:subgraphkernel:starsandpaths} we see that decreasing the number of subdivided star components from two to one makes a polynomial kernel possible. While the definition of splittable graph families that characterizes the existence of polynomial Turing kernels for \kFSubgraphTest is robust under increases by constants, this is clearly not the case for the many-one complexity of \kFSubgraphTest. \subsection{Motivation for hereditary classes} \label{section:motivate:hereditary} In this paper, we restricted our study to hereditary classes $\F$. There are a number of reasons motivating this decision. First, considering arbitrary classes $\F$ can make it very hard to prove lower bounds by polynomial-time reductions (even if the classes are decidable). For a concrete example, pick $\F$ consisting of every clique of size $2^{2^{2^i}}$ for $i\ge 1$. Then \FSubgraphTest is unlikely to be polynomial-time solvable, but this seems difficult to prove with a polynomial-time reduction as the smallest clique in~$\F$ of size exceeding~$n$ may have superpolynomial size. A difficulty of different sorts appears if $\F$ contains cliques such that the sizes of cliques in $\F$ form a more dense set of integers than in the previous example, but deciding if the clique of a particular size~$n$ is in~$\F$ takes time exponential in~$n$. These issues may be considered artifacts of trying to prove hardness by uniform polynomial-time reductions that work for every input length~$n$; potentially the issues can be avoided by formulating the complexity framework in a different way. However, there are even more substantial difficulties that appear when the class $\F$ is not hereditary. For example, let $\F$ be the set of all paths. Then $\FSubgraphTest$ is NP-hard and it does not admit a polynomial kernel, unless \containment \cite{BodlaenderDFH09}. Consider now the class $\F'$ containing, for every $i\ge 1$, the graph formed by a path of length $i$ together with $2^i$ isolated vertices. The introduction of the isolated vertices should not change the complexity of the problem, but, surprisingly, it does. The problem of finding a path of length $k$ in an $n$-vertex graph can be solved in time $2^{O(k)}\cdot n^{O(1)}$ \cite{AlonYZ95,DBLP:journals/ipl/Williams09,Bjorklund10}. Therefore, if $H$ consists of a path of length $k$ and $2^k$ isolated vertices, then these algorithms give a polynomial-time algorithm for finding $H$ in a graph $G$: the running time $2^{O(k)}\cdot n^{O(1)}$ is polynomial in the size of $H$ and $G$. Therefore, $\F'$-\SubgraphTest is polynomial-time solvable, but apparently only because finding a path of length $k$ is fixed-parameter tractable and has $2^{O(k)}n^{O(1)}$ time algorithms (note that the $2^{O(k\log k)}n^{O(1)}$ time algorithm of Monien \cite{Monien85} would not be sufficient for this argument). Therefore, it seems that we need a very tight understanding of the fixed-parameter tractability of \kFSubgraphTest to argue about its polynomial-time solvability. There are examples in the literature where the polynomial-time solvability of a problem was characterized for every (not necessarily hereditary) class $\F$ \cite{DBLP:conf/icalp/ChenTW08,DBLP:journals/tcs/DalmauJ04,1206036,380867}, but in all these results, the characterization of polynomial-time was possible only because it coincided with fixed-parameter tractability. There is certainly no such coincidence for \kFSubgraphTest (for example, finding a path of length $k$ is NP-hard, but FPT) and moreover the fixed-parameter tractability of \kFSubgraphTest is not well understood, as shown, for example, by the \textsc{Biclique} problem. All these problems disappear if we restrict $\F$ to be hereditary (e.g., adding isolated vertices certainly cannot make the problem easier and if $\F$ contains arbitrary large cliques, then $\F$ contains every clique). While this restricts the generality of our results to some extent, we believe that avoiding the difficulties discussed above more than compensates for this lack of generality. Let us also comment on the fact that, while the problems we study concern finding and packing (non-induced) subgraphs, we characterize the difficulty of such problems for each class~$\F$ of pattern graphs that is closed under induced subgraphs (i.e., for hereditary classes). The discrepancy between induced and non-induced here is entirely natural. Note that every class~$\F$ of pattern graphs that is closed under subgraphs, is also closed under induced subgraphs, and is therefore covered by our dichotomies. The fact that we can also classify~$\F$ that are merely closed under induced subgraphs, rather than normal subgraphs, gives our results extra strength. We mention in passing the classical result of Lewis and Yannakakis \cite{LewisY80} on fully characterizing the complexity of vertex-deletion problems defined by hereditary properties; these results also rely crucially on the assumption that the property is hereditary. Note that our results on \FSubgraphTest are unrelated to the results of Lewis and Yannakakis \cite{LewisY80}: their problem is related to finding induced subgraphs and the task is not to find a specific induced subgraph, but to find a subgraph belonging to the class and having a specified size. \section{Preliminaries} \label{section:preliminaries} For integers~$n$ we denote the set~$\{1, \ldots, n\}$ by~$[n]$. If~$X$ is a finite set and~$n \in \mathbb{N}$ then~$\binom{X}{n}$ is the collection of size-$n$ subsets of~$X$. Similarly, we use~$\binom{X}{\leq n}$ for the collection of all subsets of~$X$ that have size \emph{at most}~$n$, including the empty set. \subsection{Parameterized complexity and kernelization} A parameterized problem~$\Q$ is a subset of~$\Sigma^* \times \mathbb{N}$, the second component of a tuple~$(x,k) \in \Sigma^* \times \mathbb{N}$ is called the \emph{parameter}. A parameterized problem is (strongly uniformly) \emph{fixed-parameter tractable} if there exists an algorithm to decide whether $(x,k) \in \Q$ in time~$f(k)|x|^{\Oh(1)}$ where~$f$ is a computable function. A \emph{many-one kernelization algorithm} (or \emph{many-one kernel}) of size~$f \colon \mathbb{N} \to \mathbb{N}$ for a parameterized problem~$\Q \subseteq \Sigma^* \times \mathbb{N}$ is an algorithm that, on input~$(x,k) \in \Sigma^* \times \mathbb{N}$, runs in time polynomial in~$|x| + k$ and outputs an instance~$(x', k')$ with~$|x'|, k' \leq f(k)$ such that~$(x,k) \in \Q \Leftrightarrow (x', k') \in \Q$. It is a \emph{polynomial kernel} if~$f$ is a polynomial (cf.~\cite{Bodlaender09}). \begin{definition} \label{definition:turing:kernelization} Let~$\Q$ be a parameterized problem and let~$f \colon \mathbb{N} \to \mathbb{N}$ be a computable function. A \emph{Turing kernelization for~$\Q$ of size~$f$} is an algorithm that decides whether a given instance~$(x,k) \in \Sigma^* \times \mathbb{N}$ is contained in~$\Q$ in time polynomial in~$|x| + k$, when given access to an oracle that decides membership in~$\Q$ for any instance~$(x',k')$ with~$|x'|, k' \leq f(k)$ in a single step. \end{definition} We refer to a textbook~\cite{FlumG06} for more background on parameterized complexity. \subsection{Graphs} \label{section:preliminaries:graphs} All graphs we consider are finite, undirected, and simple, unless explicitly stated otherwise. A graph~$G$ consists of a vertex set~$V(G)$ and an edge set~$E(G) \subseteq \binom{V(G)}{2}$. If~$H,G$ are graphs such that~$V(H) \subseteq V(G)$ and~$E(H) \subseteq E(G)$ then~$H$ is a \emph{subgraph} of~$G$, denoted~$H \subseteq G$. For a vertex set~$X \subseteq V(G)$, the subgraph of~$G$ \emph{induced} by~$G$ is the graph with vertex set~$X$ and edge set~$E(G) \cap \binom{X}{2}$. We use~$G - X$ as a shorthand for~$G[V(G) \setminus X]$. The \emph{open neighborhood} of a vertex~$v \in V(G)$ in graph~$G$ is denoted~$N_G(v)$, while the closed neighborhood (which includes~$v$ itself) is~$N_G[v]$. If~$X$ is a vertex set then~$N_G(X) = \bigcup _{v\in X} N_G(v) \setminus X$, while~$N_G[X] = \bigcup _{v \in X} N_G[v]$. We use~$\deg_G(v)$ to denote the degree of vertex~$v$ in graph~$G$. The maximum degree of~$G$ is denoted~$\Delta(G)$. For vertex sets~$X$ and~$Y$ of a graph~$G$ we say that \emph{$X$ is universal to~$Y$} if each vertex of~$X$ is adjacent to all vertices of~$Y$. If~$G$ is a graph and~$X \subseteq V(G)$ is a vertex set, then the operation of \emph{identifying the vertices~$X$ into a single vertex} consists of removing the vertices~$X$ and their incident edges, replacing them by a single new vertex~$v_X$ whose neighborhood becomes~$N_G(X)$. If~$G$ and~$H$ are graphs, then~$G + H$ denotes the disjoint union of the two graphs. For a positive integer~$t$, we denote by~$t \cdot G$ the disjoint union of~$t$ copies of~$G$. A \emph{vertex cover} of a graph~$G$ is a set~$X \subseteq V(G)$ that contains at least one endpoint of every edge. The \emph{vertex cover number} of a graph is the size of a smallest vertex cover. By~$K_n$ ($K_{n,n}$) we denote the complete (bipartite) graph on~$n$ vertices. A complete bipartite graph is also called a \emph{biclique}. It is \emph{balanced} if its two partite classes have equal sizes. A connected bipartite graph~$G$ is $b$-thin for~$b \in \mathbb{N}$ if it has a partite class of size at most~$b$. The cycle on~$n$ vertices is denoted~$C_n$ and the path on~$n$ vertices is denoted~$P_n$. Observe that graph~$P_n$ is a path of length~$n - 1$. At various points in the paper we have to consider a graph~$G$ and the common neighbors of a vertex set~$D \subseteq V(G)$ in that graph, which is denoted~$\bigcap _{v \in D} N_G(v)$. To avoid some case distinctions, we will also allow the set~$D$ to be empty in such expressions. Since~$\bigcap _{v \in D} N_G(v)$ consists of the elements~$x$ that belong to~$N_G(v)$ for every~$v \in D$, when~$D = \emptyset$ this holds for all elements. Hence for~$D = \emptyset$ the expression~$\bigcap _{v \in D} N_G(v)$ evaluates to all elements in the universe of discourse, which will simply be~$V(G)$ unless explicitly stated otherwise. The \emph{hereditary closure} of a graph family~$\F$ is the hereditary family containing all graphs in~$\F$ and all their induced subgraphs. Let~$H$ and~$G$ be graphs and let~$P \subseteq V(H)$. A \emph{$P$-partial subgraph model} of~$H$ in~$G$ is an injection~$\phi \colon P \to V(G)$ such that for all edges~$\{u,v\} \in E(H)$ with~$\{u,v\} \subseteq P$ we have~$\{\phi(u), \phi(v)\} \in E(G)$. We say that~$P$ is the \emph{domain} of~$\phi$. For sets~$P' \subseteq P$ we will write~$\phi(P')$ to denote the set~$\{ \phi(v) \mid v \in P'\}$. A $P$-partial $H$-subgraph model is a \emph{full subgraph model} if~$P = V(H)$. If~$\phi$ is a $P$-partial subgraph model in~$G$ and~$\phi'$ is a $P'$-partial subgraph model in~$G' \subseteq G$ then~$\phi'$ is an \emph{extension} of~$\phi$ if~$P \subseteq P'$ and~$\phi(v) = \phi'(v)$ for all~$v \in P$. If~$\phi$ and~$\phi'$ are partial $H$-subgraph models with domains~$P, P' \supseteq X$ then the models \emph{agree on~$X$} if~$\phi(v) = \phi'(v)$ for all~$v \in X$. If~$\phi$ is a $P$-partial $H$-subgraph model in~$G$, then for any set~$X \subseteq V(H)$ we define the \emph{restriction of~$\phi$ to~$X$} as the partial $H$-subgraph model~$\phi|_X \colon P \cap X \to V(G)$ given by~$\phi|_X(v) = \phi(v)$ for all~$v \in P \cap X$. The restriction will sometimes be used as a partial $H[X]$-subgraph model rather than a partial $H$-subgraph model; it will be clear from the context which is meant. A \emph{separation} of a graph~$G$ is a pair~$(A, B)$ of subsets of~$V(G)$ such that~$A \cup B = V(G)$ and there are no edges between~$A \setminus B$ and~$B \setminus A$. The following observation formalizes that if~$(A,B)$ is a separation of~$H$ into two parts and we have full subgraph models for~$H[A]$ and~$H[B]$ that realize the separator~$A \cap B$ in the same way, then these models can be glued together to form a full subgraph model of~$H$. \begin{observation} \label{observation:merge:models} Let~$G$ and~$H$ be graphs, let~$(A,B)$ be a separation of~$H$, and let~$\phi$ be a partial~$H$-subgraph model in~$H$ with domain~$P \supseteq A \cap B$. If~$\phi_A$ is a full $H[A]$-subgraph model in~$G$ that extends~$\phi|_A$ and~$\phi_B$ is a full $H[B]$-subgraph model in~$G$ that extends~$\phi|_B$ such that~$\phi_A(A \setminus B) \cap \phi_B(B \setminus A) = \emptyset$, then the injection~$\phi^* \colon V(H) \to V(G)$ defined as follows: \begin{equation*} \phi^*(v) = \begin{cases} \phi_A(v) & \text{if~$v \in A \setminus B$,} \\ \phi_B(v) & \text{if~$v \in B \setminus A$,} \\ \phi_A(v) = \phi_B(v) & \text{if~$v \in A \cap B$,} \end{cases} \end{equation*} is a full $H$-subgraph model in~$G$ that extends~$\phi$. \end{observation} \subsection{Ramsey theory} We review the basic results that we need. Ramsey's Theorem states that if the edges of a sufficiently large clique are colored with a bounded number of colors, then there has to be a large \emph{monochromatic} clique with every edge having the same color. \begin{theorem}[{Cf.~\cite[Chapter 25]{GrahamGL95}}]\label{theorem:ramsey} For every choice of positive integers~$n$ and~$t$ there exists a number~$R(n,t)$ such that for every $t$-coloring~$f \colon E(K_{R(n,t)}) \to [t]$ of the edges of the complete $R(n,t)$-vertex graph, there exists a monochromatic complete subgraph with~$n$ vertices. \end{theorem} There is a variant of Ramsey's Theorem for bipartite graphs, where we are looking for a monochromatic biclique of a specific size in an edge coloring of a large biclique. \begin{theorem}[{\cite{BeinekeS76,CarnielliC1999}}]\label{theorem:bipartite:ramsey} For every choice of positive integers~$n$ and~$t$, there exists a number~$\bipRamsey(n,t)$ such that for every $t$-coloring~$f \colon E(K_{\bipRamsey(n,t),\bipRamsey(n,t)}) \to [t]$ of the edges of the balanced biclique with $\bipRamsey(n,t)$ vertices in each class, there exists a monochromatic balanced biclique with $n$ vertices in each class. \end{theorem} We also use a recent result of Atminas, Lozin, and Razgon \cite{AtminasLR12} showing that a long path implies the existence of a long induced path or a large biclique. \begin{theorem}[{\cite[Theorem 1]{AtminasLR12}}] \label{theorem:path:ramsey} For every choice of positive integers~$n$ and~$k$, there exists a number~$P_0(n,k)$ such that any graph with a path on~$P_0(n,k)$ vertices either contains an induced path on~$n$ vertices or a (not necessarily induced)~$K_{k,k}$ subgraph. \end{theorem} Note that in Theorem~\ref{theorem:path:ramsey}, the biclique $K_{k,k}$ appears as a subgraph, not as an induced subgraph. However, it is easy to strengthen Theorem~\ref{theorem:path:ramsey} in a way that we can assume that~$K_{k,k}$ is induced. \begin{corollary}\label{theorem:path:ramsey2} For every choice of positive integers~$n$ and~$k$, there exists a number~$P(n,k)$ such that any graph with a path on~$P(n,k)$ vertices either contains an induced path on~$n$ vertices or a $K_{k}$ subgraph, or an induced $K_{k,k}$ subgraph. \end{corollary} \begin{proof} Let $k'=R(k,k)$ for the function $R$ in Ramsey's Theorem (Theorem~\ref{theorem:ramsey}) and let $P(n,k)=P_0(n,k')$ for the function $P_0$ in Theorem~\ref{theorem:path:ramsey}. Then Theorem~\ref{theorem:path:ramsey} implies that every graph with at least $P(n,k)$ vertices contains either an induced path on $n$ vertices (in which case we are done) or a (not necessarily induced) $K_{k',k'}$ subgraph. Let $X$ and $Y$ be the two partite classes of the $K_{k',k'}$ subgraph. By Ramsey's Theorem, $X$ contains either a clique on $k$ vertices (in which case we are done) of an independent set $X'\subseteq X$ on $k$ vertices. Similarly, we may assume that there is an independent set $Y'\subseteq Y$ of size $k$. Now $X\cup Y$ induces a $K_{k,k}$ subgraph. \end{proof} \subsection{WK[1]-hardness proofs} \label{section:wkhardness} The complexity class \textup{WK[1]} can be defined~\cite[Section 4]{HermelinKSWW13} as the closure of \nExactSetCover under polynomial-parameter transformations, which are defined as follows. \begin{definition}[\cite{BodlaenderTY11}] \label{definition:polyParamTransform} Let~$\Q,\Q'\subseteq\Sigma^*\times \mathbb{N}$ be parameterized problems. A \emph{polynomial-parameter transformation} from~$\Q$ to~$\Q'$ is an algorithm that on input~$(x,k)\in\Sigma^*\times \mathbb{N}$ takes time polynomial in~$|x|+k$ and outputs an instance~$(x',k')\in\Sigma^*\times \mathbb{N}$ such that: \begin{itemize} \item The parameter value~$k'$ is polynomially bounded in~$k$. \item $(x',k') \in \Q'$ if and only if~$(x,k) \in \Q$. \end{itemize} \end{definition} \noindent We will use a \emph{regular} variant of the set cover problem as the starting point for our \textup{WK[1]}-hardness proofs. If~$\S$ is a set system over universe~$U$ then we will say that the pair~$(\S,U)$ has an exact cover if there is a subsystem~$\S' \subseteq \S$ such that each element of~$U$ is contained in exactly one set of~$\S'$. Recall that a set system is $r$-uniform if all sets have size exactly~$r$. \parproblemdef{\nRegularExactSetCover} {An integer~$r \geq 3$ and an $r$-uniform set system~$\S$ over a universe~$U$.} {$n := |U|$.} {Does~$(\S,U)$ have an exact cover?} \begin{lemma} \label{lemma:regularexactsetcover:wkhard} \nRegularExactSetCover is \textup{WK[1]}-hard. \end{lemma} \begin{proof} Hermelin et al.~\cite{HermelinKSWW13} proved that the variant \nExactSetCover, where sets may have different sizes, is \textup{WK[1]}-hard. The \nExactSetCover problem asks for a given set system~$\S$ over a universe~$U$ whether~$(\S,U)$ has an exact cover. The parameter is~$n := |U|$. We may assume that~$n \geq 2$, as otherwise we can solve the problem in polynomial time and output a constant-size instance that gives the same answer. We transform an instance~$(\S, U, n)$ of \nExactSetCover to an equivalent instance of \nRegularExactSetCover in two steps. First we create a system~$(\S', U')$ such that~$|U'| = 2|U|$ and~$\S$ has an exact set cover if and only if~$\S'$ has an exact set cover with exactly~$n+1$ sets. The system is built by adding~$n$ new elements~$u'_1, \ldots, u'_n$ to the universe and adding all sets~$\{ \{u'_i, \ldots, u'_j\} \mid 1 \leq i \leq j \leq n \}$ to~$\S$ to obtain the system~$\S'$. \begin{claim} $(\S,U)$ has an exact set cover if and only if~$(\S',U')$ has an exact cover consisting of~$n+1$ sets. \end{claim} \begin{claimproof} Suppose that~$(\S,U)$ has an exact set cover~$S_1, \ldots, S_k$. Observe that~$1 \leq k \leq n$, otherwise some universe element is contained in two sets. The~$n$ new elements~$u'_1, \ldots, u'_n$ that have been added to the sequence can be exactly covered with the single set~$\{u'_1, \ldots, u'_n\}$, with~$n$ sets~$\{u'_1\}, \{u'_2\}, \ldots, \{u'_n\}$, and with any number of sets between one and~$n$, since all consecutive intervals of these new elements have been added to~$\S'$. Hence we may augment the exact cover for~$(\S,U)$ with an exact cover of the new elements with~$(n+1) - k$ sets. We obtain an exact cover for~$(\S',U')$ with~$n+1$ sets. In the reverse direction, observe that all sets in~$\S'$ that contain an element from the original set~$U$, also exist in~$\S$. Hence from an exact cover of~$(\S',U')$ we can select the sets containing elements from~$U$ to obtain an exact cover for~$(\S,U)$. \end{claimproof} From the system~$(\S', U')$ we then construct another system~$(\S^*, U^*)$, as follows. Form~$U^*$ by adding~$2n^2$ new elements~$\{u^*_1, \ldots, u^*_{2n^2}\}$ to~$U'$. Define~$\S^*$ as~$\{S \cup \{u^*_i, \ldots, u^*_{i + (2n - |S|) - 1}\} \mid S \in \S' \wedge 1 \leq i \leq 2n^2 - (2n - |S|) + 1 \}$, which is~$2n$-uniform. \begin{claim} $(\S',U')$ has an exact cover consisting of~$n+1$ sets if and only if~$(\S^*, U^*)$ has an exact cover. \end{claim} \begin{claimproof} Suppose that~$(\S',U')$ has an exact cover consisting of~$n+1$ sets~$S_1, \ldots, S_{n+1}$. Observe that each set in~$\S'$ has size at most~$n$. For~$i \in [n+1]$ define~$t_i := \sum _{j=1}^{i-1} (2n-|S_j|)$, implying~$t_1 = 0$. For each~$i \in [n+1]$, define~$S'_i := S_i \cup \{u^*_{t_{i-1} + 1}, \ldots, u^*_{t_{i-1} + 2n - |S_i|}\}$. Then~$S'_i \in \S^*$ by our definition of~$\S^*$. Hence the sets~$S'_1, \ldots, S'_{n+1}$ are contained in~$\S^*$, each have size~$2n$, and are pairwise disjoint. As they contain~$2n(n+1) = 2n^2 + 2n = |U^*|$ elements in total, they form an exact set cover of~$(\S^*,U^*)$. For the reverse direction, observe that as each set in~$\S^*$ has size~$2n$ while the universe size is~$2n^2 + 2n$, any exact set cover in~$(\S^*, U^*)$ consists of exactly~$n+1$ sets~$S'_1, \ldots, S'_{n+1}$. The intersection of each set~$S'_i$ with~$U'$ is non-empty and contained in~$\S'$, by definition of~$\S^*$. Hence~$S'_1 \cap U', \ldots, S'_{n+1} \cap U'$ is an exact cover of~$(\S',U')$ consisting of~$n+1$ sets. \end{claimproof} Together the two claims show that the system~$(\S^*,U^*)$ has an exact set cover if and only if the input~$(\S,U)$ has one. System~$\S^*$ is $r$-uniform for~$r := 2n \geq 4$. The universe size~$n^* := |U^*|$ equals~$2n + 2n^2$, which is polynomial in the parameter of the input instance. Since the construction can be performed in polynomial time, it forms a valid polynomial-parameter transformation from \nExactSetCover to \nRegularExactSetCover, which concludes the proof by the \textup{WK[1]}-hardness of the former problem. \end{proof} \section{Polynomial-time solvable versus NP-complete} \label{sec:polyn-time-solv} In this section, we prove Theorem~\ref{theorem:intro:polysubgraph}, characterizing the herditary classes $\F$ for which \FSubgraphTest can be solved in randomized polynomial time. In Section~\ref{sec:poly-upper-bound}, we use the randomized matching algorithm of Mulmuley, Vazirani, and Vazirani \cite{MulmuleyVV87} (Theorem~\ref{prop:wmatching}) to solve \FSubgraphTest in randomized polynomial-time for matching-splittable hereditary families. In Section~\ref{sec:poly-lower-bound}, we prove Theorem~\ref{theorem:ramsey:matchingsplittable} characterizing matching-splittable herditary classes. In Section~\ref{subsection:poly:theorem}, we simply put together these results to complete the proof of Theorem~\ref{theorem:intro:polysubgraph}. \subsection{Upper bound} \label{sec:poly-upper-bound} \label{section:polynomialvsnpcomplete:upperbounds} Our algorithm builds on the following algebraic matching procedure. \begin{proposition}[Mulmuley, Vazirani, and Vazirani \cite{MulmuleyVV87}]\label{prop:wmatching} There exists a randomized algorithm with false negatives that, given a multigraph $G$ with nonnegative integer weights and a target weight $w_0$, checks in time polynomial in $|V(G)|+|E(G)|$ and $w_0$ whether there exists a perfect matching in $G$ of weight exactly $w_0$. \end{proposition} Note that originally the result is not stated for multigraphs, but it is easy to modify the proof accordingly. Alternatively, one can get rid of multiple edges by subdividing each edge twice and setting the weights appropriately. Let us prove Theorem~\ref{theorem:coloredmatching} using this result. \restatecoloredmatching* \begin{proof} We may assume that $C=[c]$ for some integer $i\ge 1$. Let $n=|V(G)|$ and let $s=\sum_{i=1}^{c}f(i)$ be the size of the matching we are looking for. Let us set the weight of each edge of color $i$ to $n^i$. Additionally, let us introduce a set $X$ of $n-2s$ vertices and let us connect every original vertex to every vertex of $X$ with and edge of weight 0. Let $G'$ be the resulting graph. We claim that $G$ has a matching with the required number of colors if and only $G'$ has a perfect matching of weight exactly $w_0=\sum_{i=1}^cf(i)n^i=n^{O(c)}$ (we may assume that $f(i)\le n$ otherwise there is no solution). The existence of such a perfect matching can be tested with the algorithm of Proposition~\ref{prop:wmatching} in time polynomial in $|E(G)|$ and $n^{O(c)}$. Suppose that $M$ is a matching of $G$ with $f(i)$ edges of color $i$ for every $i\in C$. Then the total weight of $M$ in $G'$ is exactly $w_0$. Let us extend $M$ the following way: if vertex $v\in V(G)$ is not covered by the matching $M$, then let us add the edge $uv$ (of weight 0) for some $u\in X$ to the matching. As exactly $n-2s=|X|$ vertices of $G$ are not covered by $M$, we may select a distinct $u\in X$ for each such edge, resulting in a perfect matching $M'$ of $G'$ having weight exactly $w_0$. Conversely, suppose that $M'$ is a perfect matching of $G'$ having weight exactly $w_0$. Ignoring the edges of weight 0, we get a matching $M$ of $G$. It is easy to see that the only way $M'$ can have weight exactly $w_0$, is if $M$ has exactly $f(i)$ edges of color $i$ for every $1\le i \le c$. Indeed, interpreting $w_0$ as a number in base-$n$ notation, this is the only way the weight of the edges in $M'$ add up to $w_0$ (note that no ``overflow'' can occur, as clearly there are at most $n/2$ edges of weight $n^i$ in $M'$). \end{proof} Equipped with Theorem~\ref{theorem:coloredmatching}, we prove Theorem~\ref{theorem:fsubgraphtest:alg}. \restatesubgraphtestalg* \begin{proof} Given a graph $H\in \F$ and arbitrary graph $G$, we proceed the following way. By assumption, there is a set $S\subseteq V(H)$ of size at most $c_\F$ such that every component of $H- S$ has at most two vertices. We can find such a set $S$ by brute force in time $|V(H)|^{O(c_\F)}\le |V(G)|^{O(c_\F)}$, which is polynomial in the input size. Let $s=2^{|S|}$ and let us fix an arbitrary bijection $\iota:[s]\to 2^{S}$ defining a numbering of the subsets of $S$. The single-vertex components of $H- S$ can be classified according to the neighborhood of the vertex in $S$: let $n^1_{i}$ be the number of components of $H- S$ where the neighborhood of the vertex is exactly $\iota(i)$. For the two-vertex components, we need to take into account the neighborhood of both vertices: for $1\le i \le j \le s$, let $n^2_{i,j}$ be the number of two-vertex components where the two vertices of the component have neighborhoods $\iota(i)$ and $\iota(j)$ in $S$, respectively. This way, we classify the two-vertex components into $s+\binom{s}{2}$ different types. The algorithm finds a subgraph model of $H$ in $G$ by considering all $S$-partial subgraph model $\phi_0$; this adds a factor of at most $|V(G)|^{|S|}\le |V(G)|^{c_\F}$ to the running time. For a fixed $\phi_0$, we need to find images for the components of $H- S$. We construct an edge-colored multigraph $G'$ the following way. Let $S^*=\phi_0(S)$. For every $v\in V(G)\setminus S^*$, we introduce vertices $v$ and $v'$ into $G'$ and add an edge $\{v,v'\}$ of color $\iota^{-1}(X)$ for every $\emptyset \subseteq X \subseteq \phi^{-1}_0(N_G(v)\cap S^*))$. If $v_1$ and $v_2$ are adjacent vertices in $G- S^*$, then we add parallel edges between $v_1$ and $v_2$ the following way. For every $\emptyset \subseteq X \subseteq \phi^{-1}_0(N_G(v_1)\cap S^*)$ and $\emptyset \subseteq Y \subseteq \phi^{-1}_0(N_G(v_2)\cap S^*)$, we let $i=\iota^{-1}(X)$, $j=\iota^{-1}(Y)$, and add an edge $\{v_1,v_2\}$ of color $(i,j)$ (if $i\le j$) or $(j,i)$ (if $j\le i$). Observe that set $C$ of colors we have used on the edges of $G$ has size $s+s+\binom{s}{2}=2^{O(c_{\F})}$. We can use the algorithm of Theorem~\ref{theorem:coloredmatching} to decide if there is a matching $M$ containing exactly $n^1_{i}$ edges of color $i$ and exactly $n^2_{i,j}$ edges of color $(i,j)$. Note that the number of colors is a fixed constant depending only on $c_{\F}$ and the size of the graph $G'$ is polynomial in $|V(G)|$ and the number of colors (as this bounds the number of parallel edges between two vertices). Therefore, the algorithm of Theorem~\ref{theorem:coloredmatching} runs in polynomial time for fixed $c_{\F}$. We claim that $\phi_0$ can be extended to a full subgraph model if and only if such a matching $M$ exists. Suppose that $\phi$ is a full subgraph model extending $\phi_0$. If vertex $u$ is a single-vertex component of $H- S$ and $u'$ is a neighbor of $u$ in $S$, then $\phi_0(u')$ is a neighbor of $\phi(u)$, or in other words, $N_H(u)\cap S\subseteq \phi^{-1}_0(N_{G}(\phi(u))\cap S^*)$. Therefore, if $v=\phi(u)$, then by construction an edge $\{v,v'\}$ of color $\iota^{-1}(N_H(v)\cap S)$ exists; let us add it to the matching $M$. If $u_1$ and $u_2$ form a two-vertex component of $H- S$ and $v_1=\phi(u_1)$, $v_2=\phi(u_2)$, then $X=N_H(u_1)\cap S$ is a subset of $\phi^{-1}_0(N_{G}(v_1)\cap S^*)$ and $Y=N_H(u_2)\cap S$ is a subset of $\phi^{-1}_0(N_{G}(v_2)\cap S^*)$. Therefore, by construction, there is an edge of color $(\iota^{-1}(X),\iota^{-1}(Y))$ (or $(\iota^{-1}(Y),\iota^{-1}(X))$) between $v_1$ and $v_2$; let us add it to the matching $M$. Observe that $M$ is indeed a matching and contains exactly $n^1_i$ edges of color $i$ and exactly $n^2_{i,j}$ edges of color $(i,j)$. For the reverse direction, suppose that $M$ contains the required number of edges from each color. In particular, there are $n^1_i$ edges of color $i$. Each edge of color $i$ is of the form $\{v,v'\}$ for some $v\in V(G)\setminus S^*$; let $V_i$ contain every such $v$. The fact that the edge $\{v,v'\}$ of color $i$ exists implies that $\iota(i)$ is a subset of $\phi^{-1}_0(N_G(v)\cap S^*)$. Therefore, if $u$ is a single-vertex component of $H- S$ with $N_H(u)=\iota(i)$, then $u$ can be mapped to any vertex $v\in V_i$, as $\phi_0(\iota(i))\subseteq N_G(v)\cap S^*$. Let us extend $\phi_0$ by mapping the $n^1_i$ such single-vertex components to $V_i$. Similarly, suppose that $u_1$ and $u_2$ form a two-vertex component of $H- S$ where the neighborhoods of the two vertices in $S$ are $\iota(i)$ and $\iota(j)$, respectively. Then $\{u_1,u_2\}$ can be mapped to any edge of $M$ with color $(i,j)$. This way, we can extend $\phi_0$ to a full subgraph model $\phi$: as $M$ is a matching, the images of distinct components of $H- S$ are disjoint. \end{proof} \subsection{Lower bound} \label{sec:poly-lower-bound} We prove Theorem~\ref{theorem:ramsey:matchingsplittable} characterizing matching-splittable graphs by a relative simple application of Ramsey's Theorem. (Recall that $P_3$ is the path on 3 vertices.) \restateramseymatching* \begin{proof} Assume for contradiction that none of the four cases holds. Then there is an integer $Q\ge 1$ such that $\F$ does not contain any of $\Gclique{Q}$, $\Gbiclique{Q}$, $Q\cdot K_3$, or $Q\cdot P_3$. Let~$r$ be the Ramsey number~$R(4Q, 2^9)$. Let~$H$ be a graph in~$\F$ that is not $3r$-matching-splittable; by the assumption that $\F$ is not matching-splittable, such an $H$ exists. Observe that if the maximum number of vertex-disjoint connected three-vertex graphs that can be packed in~$H$ is~$r$, then it is $3r$-matching-splittable: by maximality of the packing, the union of the vertices in the~$r$ graphs gives the required set of size $3r$. Therefore, if $H$ is not $3r$-matching-splittable, then there is a packing of~$r$ vertex-disjoint connected three-vertex subgraphs in~$H$. Let~$H_1, \ldots, H_r$ be such subgraphs and fix an arbitrary ordering~$v_{i,1}, v_{i,2}, v_{i,3}$ of the three vertices in each subgraph~$H_i$. If we consider two subgraphs~$H_i$ and~$H_{i'}$, then the adjacencies between the three vertices of~$H_i$ and the three vertices of~$H_{i'}$ are characterized exactly by the $3 \times 3$ adjacency matrix whose rows correspond to~$V(H_i)$ and whose columns correspond to~$V(H_j)$, with a one in cells corresponding to adjacent vertex pairs and a zero in the remaining cells. Since there are~$2^{3 \cdot 3} = 2^9$ different incidence matrices, the number of distinct ways in which two subgraphs~$H_i, H_j$ can be adjacent (under the chosen vertex ordering) is~$2^9$. We use these $3\times 3$ matrices to create an auxiliary graph~$F$ as follows: it is an $r$-vertex complete graph whose vertices are in correspondence with the subgraphs~$H_1, \ldots, H_r$. Number the~$2^9$ possible $3 \times 3$ adjacency matrices arbitrarily from~$1$ to~$2^9$ and give an edge~$\{i,j\} \in E(F)$ the color corresponding to the adjacency between~$H_i$ and~$H_j$. Since~$F$ has~$r = R(4Q, 2^9)$ vertices and its edges have been colored with~$2^9$ distinct colors, there is a monochromatic complete subgraph on~$4Q$ vertices. This subgraph of~$F$ corresponds to~$4Q$ subgraphs in the list~$H_1, \ldots, H_r$ that pairwise all have the same adjacencies to each other. Since there are only two different connected three-vertex graphs ($K_3$ and~$P_3$) this implies there are at least~$2Q$ pairwise vertex-disjoint, isomorphic subgraphs of~$H$ that all have the same adjacency to each other; denote the indices of these subgraphs by~$\I$. Suppose that the adjacency-type corresponding to the color of the monochromatic subgraph has at least one edge, i.e., that there are indices~$a,b \in [3]$ such that~$v_{i,a}$ is adjacent to~$v_{i',b}$ for all~$i \neq i' \in \I$. If this occurs for~$a = b$, then the set~$\{ v_{i,a} \mid a \in \I \}$ is a clique of size~$|\I| \geq 2Q$ in~$H$; but then, since~$\F$ is hereditary, family~\F contains the $\Gclique{2Q}$, which contradicts our choice of~$Q$. Now suppose that $a\neq b$ holds. Let $\I_1$ and $\I_2$ be two disjoint subsets of $\I$, each of size $Q$, and consider the vertex set~$\{v_{i,a} \mid i \in \I_1\} \cup \{v_{i,b} \mid i \in \I_2\}$. Since there are no edges between~$v_{i,a}$ and~$v_{i', a}$ for~$i,i' \in \I$ (by assumption that the previous case~$a=b$ does not apply), and similarly there are no edges between~$v_{i,b}$ and~$v_{i',b}$ for~$i,i' \in \I$, while all edges between~$v_{i,a}$ and~$v_{i',b}$ are present for~$i \in \I_1$ and~$i' \in \I_2$, the defined vertex set of size~$2Q$ induces a balanced biclique in~$H$. But then~\F contains $\Gbiclique{Q}$ (since~\F is hereditary), again contradicting our choice of~$Q$. We may therefore conclude that vertices in different subgraphs~$H_i, H_{i'}$ with~$i \neq i' \in \I$ are not adjacent to each other. But then the vertices in the graphs~$H_i$ for~$i \in \I$ induce~$2Q$ disjoint copies of the same three-vertex graph, that is, either $Q\cdot K_3$ or $Q\cdot P_3$ is an induced subgraph of $H$, contradicting the choice of~$Q$. \end{proof} \subsection{Proof of the dichotomy for polynomial-time solvability of subgraph problems} \label{subsection:poly:theorem} By combining the algorithm of Section~\ref{sec:poly-upper-bound} and the characterization proved in Section~\ref{sec:poly-lower-bound}, the proof of Theorem~\ref{theorem:intro:polysubgraph} follows. \restateintropolysubgraph* \begin{proof} Let $\F$ be a hereditary class of graphs. If $\F$ is matching-splittable, then Theorem~\ref{theorem:fsubgraphtest:alg} shows that \FSubgraphTest can be solved in randomized polynomial time. If $\F$ is not matching-splittable, then $\F$ is the superset of one of the four classes listed in Theorem~\ref{theorem:ramsey:matchingsplittable}. If $\F$ is a superset of $\Fclique$, then the NP-hard \textsc{Clique} problem can be reduced to \FSubgraphTest and hence \FSubgraphTest is also NP-hard. Similarly, if $\F$ is a superset of $\Fbiclique$, then \textsc{Biclique} can be reduced to \FSubgraphTest. Suppose now that $\F$ contains $n\cdot K_3$ for every $n\ge 1$. Then $K_3$-\Packing, which is NP-hard by Theorem~\ref{theorem:fpacking:pvsnp}, can be reduced to \FSubgraphTest: an instance $(G,K_3,t)$ of $K_3$-\Packing can be expressed as an instance $(G,t\cdot K_3)$ of \FSubgraphTest. The situation is similar if $\F$ contains $n\cdot P_3$ for every $n\ge 1$; note that $P_3$-\Packing is also NP-hard by Theorem~\ref{theorem:fpacking:pvsnp}. Therefore, we have shown that if $\F$ is not matching-splittable, then \FSubgraphTest is NP-hard, completing the proof Theorem~\ref{theorem:intro:polysubgraph}. \end{proof} \section{Computing representative sets for subgraph detection} \label{sec:comp-repr-sets} In this section, we present the main technology behind the kernelization results of the paper: a marking algorithm that can be used to find a representative set of small and thin bipartite subgraphs that is sufficient for the solution. This marking algorithm is used in the positive side of both Theorem~\ref{theorem:intro:packing} for small/thin classes (Section~\ref{subsection:packing:upperbounds}) and in the positive side of Theorem~\ref{theorem:intro:turingsubgraph} for splittable classes (Section~\ref{subsection:subgraph:turing:upperbounds}). The application for Theorem~\ref{theorem:intro:turingsubgraph} is more general, as it involves the more general splittable proprety. We present the results in a way suitable for this more general application. While this is more general than what is need for Theorem~\ref{theorem:intro:packing}, it makes no sense to present two versions of essentially the same algorithm. \subsection{Finding thin bipartite subgraphs} The definition of small/thin and splittable graphs involves components of bounded size and thin bipartite graphs. Therefore, the very least, we should be able to find such components efficiently. If a component has at most $a$ vertices, then we can try all $n^a$ possible images in $G$ by brute force. If a component is $b$-thin bipartite, then we can try all $n^b$ possible images for the partite class containing at most $b$ vertices and then we have to solve a bipartite matching problem to find the location of the vertices in the larger class. The follow lemma presents this reduction to matching in the slightly more general context when the images of a set of vertices, including every vertex in the smaller class, are already fixed. \begin{lemma} \label{lemma:compute:extension:partial:specifiedbipartite:subgraph} Let~$G$ be a graph, let~$H$ be a bipartite graph with partite sets~$A = \{a_1, \ldots, a_\alpha\}$ and~$B = \{b_1, \ldots, b_\beta\}$, and let~$\phi$ be a $P$-partial subgraph model of~$H$ with~$A \subseteq P \subseteq V(H)$. One can compute a full subgraph model of~$H$ in~$G$ that extends~$\phi$, or determine that no such model exists, in polynomial time by computing a maximum matching in a bipartite graph of order at most~$2|V(G)|$. \end{lemma} \begin{proof} If~$|V(H)| > |V(G)|$ then obviously there is no full model of~$H$ in~$G$ and we output \no. Otherwise we proceed as follows. Let~$\N := \{N_H(b) \mid b \in B \setminus P\}$ be the set of different neighborhoods in~$A$ that have to be realized for the vertices that are not yet specified in the partial model. Since~$H$ is bipartite we have~$A' \subseteq A$ for all~$A' \in \N$. Number the sets in~$\N$ as~$\N = \{A'_1, \ldots, A'_t\}$. For each~$i \in [t]$ let~$n(i)$ be the number of vertices in~$B \setminus P$ whose neighborhood in~$H$ is exactly~$A'_i$. As~$A \subseteq P$, it follows that there is an extension of~$\phi$ to a full model if and only if we can assign to each set~$A'_i$ a set~$B'_i \subseteq \bigcap _{a \in A'_i} N_G(\phi(a)) \setminus \phi(P)$ of size~$n(i)$ such that the sets~$B'_i$ are pairwise disjoint. The crucial insight is that this condition can be checked by computing a matching in a related bipartite graph~$\hat{G}$ that is defined as follows. The partite set~$\hat{B}$ of~$\hat{G}$ consists of the vertices~$V(G) \setminus \phi(P)$. The other partite set~$\hat{A}$ contains, for each set~$A'_i$ with~$i \in [t]$, exactly~$n(i)$ vertices~$w_{i,1}, \ldots, w_{i, n(i)}$ that are false twins in~$\hat{G}$. Each vertex~$w_{i,j}$ for~$j \in [n(i)]$ is adjacent in~$\hat{G}$ to~$\bigcap _{a \in A'_i} N_G(\phi(a)) \setminus \phi(P)$. Since~$|V(H)| \leq |V(G)|$, the order of~$\hat{G}$ is at most~$2|V(G)|$. \begin{claim} If there is a maximum matching in~$\hat{G}$ that saturates~$\hat{A}$, then~$\phi$ can be extended to a full $H$-subgraph model in~$G$: for each~$i \in [t]$, let~$v_{i,1}, \ldots, v_{i, n(i)} \subseteq V(H) \setminus P$ be the vertices whose $H$-neighborhood is~$A'_i$ and set~$\phi(v_{i,j})$ to the matching partner of~$w_{i,j}$. If a maximum matching in~$\hat{G}$ has size less than~$|\hat{A}|$ then there is no extension of~$\phi$ to a full subgraph model of~$H$ in~$G$. \end{claim} \begin{claimproof} Suppose that~$M$ is a matching in~$\hat{G}$ saturating~$\hat{A}$. Consider the extension of~$\phi$ suggested in the claim. Then a vertex~$v_{i,j}$ is mapped to the matching partner of~$w_{i,j}$, which is a neighbor of~$w_{i,j}$ in~$\hat{G}$. As the neighbors of~$w_{i,j}$ in~$\hat{G}$ are exactly the vertices in~$\bigcap _{a \in A'_i} N_G(\phi(a)) \setminus \phi(P)$, this maps~$v_{i,j}$ to a vertex of~$G$ not yet used in the model that is adjacent to the $\phi$-model of all of~$v_{i,j}$'s neighbors in~$H$. As the matching ensures that all images assigned in this way are distinct, we obtain a valid $H$-subgraph model in~$G$. We prove the second statement by contraposition: if there is an extension of~$\phi$ to a full subgraph model of~$H$, then there is a matching in~$\hat{G}$ saturating~$\hat{A}$. Suppose that~$\phi^*$ is a full $H$-subgraph model that extends~$\phi$. For each~$i \in [t]$ let~$v_{i,1}, \ldots, v_{i,n(i)}$ be as in the statement of the claim. Match~$w_{i,j} \in \hat{A}$ to~$\phi(v_{i,j}) \in \hat{B}$ for all~$i$ and~$j$ to obtain a matching saturating~$\hat{A}$, which has size~$|\hat{A}|$. This proves the claim. \end{claimproof} The claim shows how to construct a $H$-model that extends~$\phi$, if one exists. Since the construction of~$\hat{G}$ can be done in time polynomial in~$|V(G)|$ and bipartite matching is polynomial-time, for example using the Hopcroft-Karp algorithm~\cite{HopcroftK73}, this concludes the proof of Lemma~\ref{lemma:compute:extension:partial:specifiedbipartite:subgraph}. \end{proof} We now present the algorithm for finding a $b$-thin bipartite graph. The following lemma formulates this in a more general way suitable for use in $(a,b,c,d)$-splittable graphs: there is a set $D$ such that $H-D$ is a thin bipartite graph, and all but a bounded number of vertices in $H-D$ are universal to $D$. \begin{lemma} \label{lemma:compute:extension:partial:separated:subgraph} Let~$G$ be a graph, let~$H$ be a graph with a (possibly empty) vertex set~$D \subseteq V(G)$ such that~$H' := H - D$ is a bipartite graph with partite sets~$A = \{a_1, \ldots, a_\alpha\}$ and~$B = \{b_1, \ldots, b_\beta\}$, and let~$\phi_0$ be a $P_0$-partial subgraph model of~$H$ in~$G$ with~$D \subseteq P_0$. Let~$B_N \subseteq B$ contain the vertices whose closed neighborhood in~$H$ is not universal to~$D$. One can compute a full subgraph model of~$H$ in~$G$ that extends~$\phi_0$, or determine that no such model exists, in time~$|V(G)|^{\Oh(1 + |(A \cup B_N) \setminus P_0|)}$. \end{lemma} \begin{proof} We consider all~$\Oh(|V(G)|^{|A \cup B_N \setminus P_0|})$ possible ways to map the unspecified vertices~$(A \cup B_N) \setminus P_0$ to distinct vertices of~$V(G) \setminus P_0$. For each resulting $P'$-partial $H$-subgraph model~$\phi'$ we test whether it is valid, i.e., whether for all edges~$\{a,b\} \in E(H)$ with~$a, b \in P'$ we have~$\{\phi'(a), \phi'(b)\} \in E(G)$. If this is the case, then we want to invoke Lemma~\ref{lemma:compute:extension:partial:specifiedbipartite:subgraph} to determine whether~$\phi'$ can be extended to a full model of~$H$ that extends~$\phi'$ (and therefore~$\phi$). The crucial observation is that such an extension of~$\phi'$ exists if and only if the partial $H'$-subgraph model~$\phi'|_{V(H')}$ can be extended to a full model of~$H'$ in the graph~$G[\phi'(V(H')) \cup \bigcap _{v \in D} N_G(\phi(v))]$. This follows from the fact that all vertices of~$H'$ that are not assigned an image by~$\phi'$ are contained in~$B \setminus B_N$, are therefore universal in~$H$ to~$D$, and must therefore be mapped to members of~$\bigcap _{v \in D} N_G(\phi(v))$ by any extension. Hence by invoking Lemma~\ref{lemma:compute:extension:partial:specifiedbipartite:subgraph} we can test whether a particular choice of~$\phi'$ can be extended to a full model of~$H$ in~$G$ that extends~$\phi_0$. If any valid extension~$\phi'$ of~$\phi_0$ results in a full model then the first such model is given as the output. If Lemma~\ref{lemma:compute:extension:partial:specifiedbipartite:subgraph} never returns an extension then, since we try all possibilities for mapping~$(A \cup B_N) \setminus P_0$ to the free vertices~$V(G) \setminus \phi_0(P_0)$ in~$G$, no full model extending~$\phi_0$ exists and we output \no. The time bound follows from the fact that we try~$\Oh(|V(G)|^{|(A \cup B_N) \setminus P_0|})$ possibilities that can each be tested for feasibility in polynomial time. \end{proof} \subsection{Representative sets for small separators into unbalanced bipartite graphs} \label{sec:repr-sets-thin} This section contains the main technical part of the kernelization algorithms: the marking algorithm for thin bipartite graphs. We present it in a way suitable for $(a,b,c,d)$-splittable graphs: there is a set $D$ such that $H-D$ is a thin bipartite graph, and all but a bounded number of vertices in $H-D$ are universal to $D$. \begin{lemma} \label{lemma:compute:representative:set:subgraph} There is an algorithm with the following specifications. The input is a graph~$G$, a graph~$H$ with a (possibly empty) vertex set~$D \subseteq V(H)$ such that~$H' := H - D$ is a connected bipartite graph with partite sets~$A = \{a_1, \ldots, a_\alpha\} \neq \emptyset$ and~$B = \{b_1, \ldots, b_\beta\}$, an integer~$\ell$, and a partial subgraph model~$\phi_0$ of~$H$ with domain~$P_0 \supseteq D$. Define~$B_U := \{b \in B \mid \bigcap_{v \in N_{H'}[b]} N_H(v) \supseteq D\}$ and let~$B_N := B \setminus B_U$. Let~$h := |V(H)|$. The output is a set~$X \subseteq V(G)$ with the following properties. \begin{enumerate}[(P1)] \item $|X| \leq (h^2 \ell + h^3)^{\mu+1}(1 + 2^{\alpha}(\ell + h))$, where~$\mu := |(A \cup B_N) \setminus P_0| + \sum _{a \in A} \max (0, \alpha - |N_H(a) \cap B_U \cap P_0|)$. \label{rep:set:subgr:size} \item For any vertex set~$Z \subseteq V(G)$ of size at most~$\ell$, if~$G - Z$ contains a full subgraph model~$\phi_1$ of~$H$ that extends~$\phi_0$, then~$G[X] - Z$ contains a full subgraph model~$\phi'_1$ of~$H$ with~$\phi'_1(v) = \phi_0(v)$ for all~$v \in D$.\label{rep:set:subgr:preservesmodel} \end{enumerate} The running time of the algorithm is polynomial in~$2^\alpha + |V(G)|^\mu$. \end{lemma} \begin{proof} The algorithm is recursive and branches in a bounded number of directions to explore different ways in which the partial subgraph model~$\phi$ can be extended to a full model. The depth of the recursion is bounded by the measure~$\mu$, which will decrease in each recursive call. Throughout the proof we use the convention (see Section~\ref{section:preliminaries:graphs}) that if~$D = \emptyset$, then the common neighborhood of~$D$ in~$G$ equals~$V(G)$. We consider several cases. These correspond to the cases in the proof of Lemma~\ref{lemma:bicliquemarking:specialcase:general}, except that the first case in the lemma is not present here because it collapses into Case 2.b. \textbf{Case 1: There is a vertex~$a \in A \setminus P_0$ with~$|N_H(a) \cap P_0 \cap B_U| \geq \alpha$.} In this case there is a vertex~$a$ whose model is not yet specified by~$\phi_0$, but for which many neighbors in the $B$-side of the bipartite graph~$H'$ already have been assigned an image. If there are many options for valid images for~$a$ (Case 1.a) then this implies the existence of a large biclique, which we can mark to preserve models of the bipartite graph~$H'$. If there are few options for valid images for~$a$ (Case 1.b) then we can branch into all possible options, recursively calculate a representative set, and output the union of these sets. Formally, we make another distinction. Let~$R := N_H(a) \cap P_0 \cap B_U$. \textbf{Case 1.a: The set~$\phi(R \cup D)$ has at least~$\ell + h$ common neighbors in~$G$.} If the number of common neighbors~$\bigcap _{v \in R \cup D} N_G(\phi(v))$ is at least~$\ell + h$, then there is a large biclique in the common neighborhood of~$\phi_0(D)$ in~$G$: one side of the biclique is~$\phi(R)$ and the other side is formed by the common neighbors of~$\phi(R \cup D)$. Let~$X$ contain~$\phi(P_0)$ together with~$\ell + h$ common $G$-neighbors of~$\phi(R \cup D)$. It is easy to see that~\ref{rep:set:subgr:size} is satisfied. To see that~\ref{rep:set:subgr:preservesmodel} holds, observe the following. If~$\phi_1$ is a full $H$-model in~$G - Z$ that extends~$\phi_0$, then~$\phi(R \cup D)$ is disjoint from~$Z$. Build a full $H$-model~$\phi'_1$ as follows. Let~$\phi'_1(d) = \phi_0(d)$ for all~$d \in D$. Map~$A$ to~$\alpha$ arbitrary vertices in~$R$. Map~$B$ to vertices in the set~$\phi(R \cup D) \setminus Z$, which has size at least~$h - \alpha \geq \beta$. It is easy to verify that~$\phi'_1$ is a full $H$-model that coincides with~$\phi_0$ on the vertices of~$D$. In particular, all vertices~$\phi_1(V(H'))$ lie in the common neighborhood of~$\phi_1(D) = \phi_0(D)$, while~$G[\phi_1(H')]$ is a biclique with partite sets of sizes~$\alpha$ and~$\beta$. \textbf{Case 1.b: The set~$\phi(R \cup D)$ has fewer than~$\ell + h$ common neighbors in~$G$.} Observe that, to extend~$\phi_0$ to a full subgraph model~$\phi_1$ of~$H$, the image of vertex~$a$ has to be adjacent in~$G$ to the images of all vertices in~$N_H(a) \cap P_0$. Hence~$\phi_1(a) \in \bigcap _{v \in R} N_G(\phi_0(v))$ for all valid extensions~$\phi_1$ of~$\phi_0$. Since~$|N_H(a) \cap P_0 \cap B_U| = |R| \geq \alpha \geq 1$ by the assumption of Case 1, the vertex~$a$ has a neighbor in~$H$ in the set~$B_U$. Since~$B_U$ contains the vertices whose \emph{closed neighborhood} in~$H$ is universal to~$D$, this implies that~$a$ is universal to~$D$ in~$H$. Hence in any valid extension~$\phi_1$ of~$\phi_0$ to a full $H$-subgraph model in~$G$, the image of~$a$ has to be contained in~$\bigcap _{v \in D} N_G(\phi_0(v))$. Hence~$\phi_1(a) \in T := \bigcap _{v \in R \cup D} N_G(\phi_0(v))$. By the assumption of Case 1.b we have~$|T| < \ell + h$ and hence there are only few options for valid images of~$a$. We now do the following. Initialize~$X$ as an empty set. For each~$a' \in T$ we extend the partial subgraph model~$\phi_0$ to~$\phi_0^{a \mapsto a'}$ by setting~$\phi_0^{a \mapsto a'}(a) := a'$. We then recursively invoke the algorithm for the model~$\phi_0^{a \mapsto a'}$ and add the resulting set~$X^{a \mapsto a'}$ to~$X$. Since each such extension assigns strictly more $A$-vertices an image than~$\phi_0$, the measure~$\mu^{a \mapsto a'}$ of each recursive call is strictly smaller than the measure~$\mu$ for the current invocation. We can therefore bound the size of~$X$ as follows: \begin{align*} |X| &\leq \sum _{a' \in T} |X^{a \mapsto a'}| \leq \sum _{a' \in T} (h^2 \ell + h^3)^{(\mu-1)+1}(1 + 2^{\alpha}(\ell + h)) \leq |T| (h^2 \ell + h^3)^{(\mu-1)+1}(1 + 2^{\alpha}(\ell + h)) \\ &\leq (\ell + h) (h^2 \ell + h^3)^{(\mu - 1)+1}(1 + 2^{\alpha}(\ell + h)) \leq (h^2 \ell + h^3)^{\mu+1}(1 + 2^{\alpha}(\ell + h)), \end{align*} which satisfies~\ref{rep:set:subgr:size}. To see that~\ref{rep:set:subgr:preservesmodel} is satisfied, consider an arbitrary set~$Z \subseteq V(G)$ of size at most~$\ell$. If~$\phi_1$ is a full $H$-model in~$G - Z$ that extends~$\phi_0$, then by the observation above we have~$a' := \phi_1(a) \in T$. Hence~$\phi_1$ is a full $H$-model in~$G - Z$ that extends~$\phi_0^{a \mapsto a'}$. By the guarantee of the recursive call this implies that~$G[X^{a \mapsto a'}] - Z$ contains a full $H$-subgraph model~$\phi'_1$ with~$\phi'_1(v) = \phi_0(v)$ for all~$v \in D$. As~$X^{a \mapsto a'} \subseteq X$ this model also exists in~$G[X] - Z$, proving~\ref{rep:set:subgr:preservesmodel}. \textbf{Case 2: All~$a \in A \setminus P_0$ satisfy~$|N_H(a) \cap P_0 \cap B_U| < \alpha$.} Let~$\hat{B}_U := \{b \in B_U \mid N_H(b) \subseteq P_0\}$. Intuitively, we can easily identify a small set of vertices to add to~$X$ in order to preserve images for the vertices of~$\hat{B}_U$, because the images of all $H$-neighbors of vertices in~$\hat{B}_U$ are already specified. On the other hand, one can verify that specifying the image of a vertex in~$V(H) \setminus (P_0 \cup \hat{B}_U)$ decreases the measure~$\mu$, which allows us to deal with those vertices by branching. We proceed as follows. Let~$\hat{H} := H - \hat{B}_U$ and let~$\hat{\phi}_0$ be the partial model~$\phi_0$ restricted to the vertices of~$\hat{H}$; we interpret it as a partial~$\hat{H}$-subgraph model. By repeated application of Lemma~\ref{lemma:compute:extension:partial:separated:subgraph} we can compute a maximal packing~$\Phi$ of partial~$H - \hat{B}_U$-subgraph models in~$G$ that extend~$\hat{\phi}_0$, such that for distinct~$\phi, \phi' \in \Phi$ we have~$\phi(V(\hat{H}) \setminus P_0) \cap \phi'(V(\hat{H}) \setminus P_0) = \emptyset$. Informally, this says that the models in~$\Phi$ extend~$\hat{\phi}_0$ onto disjoint sets of vertices. (If~$V(\hat{H}) \subseteq P_0$ then there is nothing left to extend, and we define~$\Phi$ as a set containing~$\ell + h$ copies of~$\hat{\phi}_0$ to trigger Case 2.b.) The size of the computed packing~$\Phi$ of full~$\hat{H}$-subgraph models is the basis for the last case distinction. \textbf{Case 2.a: $|\Phi| < \ell + h$.} If the packing~$\Phi$ contains less than~$\ell + h$ models of~$\hat{H}$ in~$G$ that extend~$\hat{\phi}_0$ to disjoint sets of vertices, then this is a \emph{maximal} packing and we can use it to guide a branching step into a bounded number of relevant extensions of~$\phi_0$. Define the set~$S := \bigcup _{\phi \in \Phi} \phi(V(\hat{H}) \setminus P_0)$. Since~$|\Phi| < \ell + h$, the size of~$S$ is bounded by~$(\ell + h)\cdot h$. Every $H$-subgraph model~$\phi_1$ that extends~$\phi_0$ contains a $\hat{H}$-subgraph model~$\hat{\phi}_1$ that extends~$\hat{\phi}_0$. By the maximality of~$\Phi$, such a model~$\hat{\phi}_1$ maps at least one vertex to~$S$. Since the extension~$\hat{\phi}_1$ maps all vertices of~$P_0$ in the same way as~$\hat{\phi}_0$, we find that all~$H$-subgraph models~$\phi_1$ that extend~$\phi_0$ map at least one vertex of~$V(\hat{H}) \setminus P_0$ to a vertex in~$S$; we will branch on all possibilities, as follows. For each vertex~$v \in V(\hat{H}) \setminus P_0$, for each vertex~$v' \in S$, we extend~$\phi_0$ to~$\phi_0^{v \mapsto v'}$ by setting~$\phi_0^{v \mapsto v'}(v) := v'$. If this results in a valid partial $H$-subgraph model (i.e., if~$v'$ is adjacent in~$G$ to all vertices in~$\phi_0(N_H(v) \cap P_0)$) then we recursively invoke the algorithm to compute a set~$X^{v \mapsto v'}$ that preserves $H$-subgraph models that extend~$\phi_0^{v \mapsto v'}$. We let~$X$ be the union of the resulting sets. To bound the size of~$X$ we have to consider the measure associated with the recursive calls. Consider a recursive call for~$\phi_0^{v \mapsto v'}$, let~$P_0^{v \mapsto v'}$ be the domain of the partial model, and let~$\mu^{v \mapsto v'}$ be the measure of the recursive call. If~$v \in (A \cup B_N) \setminus P_0$ then~$|(A \cup B_N) \setminus P_0| > |(A \cup B_N) \setminus P_0^{v \mapsto v'}|$ since~$v \in ((A \cup B_N) \cap P_0^{v \mapsto v'}) \setminus P_0$. It then easily follows from the definition of measure that~$\mu > \mu^{v \mapsto v'}$. Let us now argue that in the other cases, the measure of the recursive calls is also strictly smaller than~$\mu$. If~$v \in (V(\hat{H}) \setminus P_0) \setminus (A \cup B_N)$ then, as~$\hat{H}$ does not contain the vertices~$\hat{B}$ and~$D \subseteq P_0$, it follows that~$v \in B_U \setminus \hat{B}$. By the definition of~$\hat{B}$ this implies that~$N_H(v)$ is not fully contained in~$P_0$. Since~$D \subseteq P_0$ by the precondition to the lemma, while~$v$ is contained in the~$B$-side of the bipartite graph~$H' = H - D$, this implies that there is a vertex~$a \in (A \cap N_H(v)) \setminus P_0$, i.e., that at least one $H$-neighbor of~$v$ is contained in the $A$-side of the bipartite graph and no image is specified for it by~$\phi_0$. By the precondition to Case 2 this implies that~$|N_H(a) \cap P_0 \cap B_U| < \alpha$. Let~$P_0^{v \mapsto v'}$ be the domain of~$\phi_0^{v \mapsto v'}$. Then the previous argumentation shows that~$\max (0, \alpha - |N_H(a)\cap B_U \cap P_0|) > \max (0, \alpha - |N_H(a) \cap B_U \cap P^{v \mapsto v'}_0|)$, which shows that the term for~$a$ in the measure of the recursive call to~$\phi^{v \mapsto v'}_0$ is strictly smaller than for the current call. It is easy to verify that the other terms in the measure do not increase. Hence the measure associated to all recursive calls is less than~$\mu$, guaranteeing that each recursive call outputs a set~$X^{v \mapsto v'}$ of size at most~$(h^2 \ell + h^3)^{(\mu-1)+1}(1 + 2^{\alpha}(\ell + h))$. This allows us to bound the size of~$X$ as follows: \begin{align*} |X| &\leq \sum _{v \in V(\hat{H}) \setminus P_0} \sum _{v' \in S} |X^{v \mapsto v'}| \leq \sum _{v \in V(\hat{H}) \setminus P_0} \sum _{v' \in S} (h^2 \ell + h^3)^{(\mu-1)+1}(1 + 2^{\alpha}(\ell + h)) \\ &\leq |V(\hat{H}) \setminus P_0| \cdot |S| \cdot (h^2 \ell + h^3)^{\mu}(1 + 2^{\alpha}(\ell + h)) \\ &\leq [h] \cdot [(\ell + h)h] \cdot (h^2 \ell + h^3)^{\mu}(1 + 2^{\alpha}(\ell + h)) \\ &\leq (h^2 \ell + h^3) (h^2 \ell + h^3)^{\mu}(1 + 2^{\alpha}(\ell + h)) \leq (h^2 \ell + h^3)^{\mu+1}(1 + 2^{\alpha}(\ell + h)). \end{align*} Hence~\ref{rep:set:subgr:size} is satisfied. To see that~\ref{rep:set:subgr:preservesmodel} is also satisfied, consider some set~$Z \subseteq V(G)$ and assume that~$G - Z$ contains a full $H$-subgraph model~$\phi_1$ that extends~$\phi_0$. As argued above,~$\phi_1$ maps at least one vertex of~$v \in V(\hat{H}) \setminus P_0$ to a vertex~$v' \in S$. Hence these is a choice of~$v$ and~$v'$ such that~$\phi_1$ is a full $H$-subgraph model in~$G - Z$ that extends a model~$\phi_0^{v \mapsto v'}$ on which we recursed. By the correctness guarantee for the recursive call, this implies that there is a full $H$-subgraph model~$\phi'_1$ in~$G[X^{v \mapsto v'}] - Z$ such that~$\phi'_1(u) = \phi_0(u)$ for all~$u \in D$. As~$X^{v \mapsto v'} \subseteq X$ this model also exists in~$G[X] - Z$, which shows that~\ref{rep:set:subgr:preservesmodel} is satisfied. \textbf{Case 2.b: $|\Phi| \geq \ell + h$.} In the last case the set~$\Phi$ contains at least~$\ell + h$ partial $\hat{H}$-subgraph models in~$G$ that extend~$\phi_0$ onto pairwise disjoint sets. Recall that these partial models are trivial (they coincide with~$\phi_0$) in the case that~$V(\hat{H}) \subseteq P_0$. Let~$\Phi' \subseteq \Phi$ contain exactly~$\ell + h$ disjoint extensions of~$\hat{\phi}_0$ to full models of~$\hat{H}$. We initialize~$X$ as~$\bigcup _{\phi \in \Phi} \phi(V(\hat{H}))$. For each subset~$A' \subseteq A \cap P_0$, we add~$\ell + h$ vertices from the set~$\bigcap _{v \in A' \cup D} N_G(\phi_0(v))$ to~$X$ (or all such vertices, if the number of common neighbors is less than~$\ell + h$). The result is output as the set~$X$. It is easy to see that~$|X| \leq (\ell + h)h + 2^{|A \cap P_0|}(\ell + h) \leq (h \ell + h^2) + 2^{\alpha}(\ell + h)$, which satisfies~\ref{rep:set:subgr:size} since~$\mu \geq 0$. It remains to argue that~\ref{rep:set:subgr:preservesmodel} holds. Let~$Z \subseteq V(G)$ be a set of size at most~$\ell$ and assume that~$G - Z$ contains a full $H$-subgraph model~$\phi_1$ that extends~$\phi_0$. We build a full $H$-subgraph model~$\phi'_1$ in~$G[X] - Z$, as follows. \begin{itemize} \item For all vertices~$v \in (A \cup D) \cap P_0$ we set~$\phi'_1(v) := \phi_0(v) = \phi_1(v)$; the last equality holds since~$\phi_1$ extends~$\phi_0$. As~$D \subseteq P_0$ by the precondition to the lemma, it remains to define an image for the vertices in~$A \setminus P_0$ and for the vertices in~$B$. As~$\phi_0$ is a partial subgraph model of~$H$, whenever there is an edge in~$H$ between vertices of~$(A \cup D) \cap P_0$, these assignments ensure there is an edge between the images of the vertices as well. Since the image of~$\phi_1$ is disjoint from~$Z$, no vertex is mapped to a member of~$Z$ in this step. As~$\Phi$ is not empty and for any~$\phi \in \Phi$ the set~$\phi((A \cup D) \cap P_0) = \phi_1((A \cup D) \cap P_0)$ is contained in~$X$, the image of the partial model~$\phi'_1$ constructed so far lies in~$G[X] - Z$. \item For all vertices~$b \in \hat{B}$, do the following. If~$\phi_1(b) \in X$ then let~$\phi'_1(b) := \phi_1(b)$. Now consider what happens if~$\phi_1(b) \not \in X$. By definition of~$\hat{B}$ we have~$N_H(b) \subseteq P_0$. As~$\hat{B} \subseteq B_U$ we know by definition of the latter set that~$D \subseteq N_H(b)$. Since~$H' = H - D$ is bipartite and~$b$ is contained in its $B$-side, there is a set~$A' \subseteq A$ such that~$N_H(b) = D \cup A'$. As~$\phi_1$ forms a valid $H$-subgraph model that extends~$\phi_0$ we have~$\phi_1(b) \in \bigcap _{v \in N_H(b)} \phi_1(v) = \bigcap _{v \in D \cup A'} \phi_0(v)$. While constructing the set~$X$ we have considered the set~$A' \subseteq A \cap P_0$ and have added common neighbors of~$\phi_0(A' \cup D)$ to~$X$. Since~$\phi_1(b) \not \in X$, it follows that the number of common $G$-neighbors of~$\phi_0(A' \cup D)$ exceeded~$\ell + h$ and~$\phi_1(b)$ was not chosen to be added to~$X$. This implies that~$(X \cap \bigcap _{v \in A' \cup D} N_G(\phi_0(v))) \setminus Z$ has size at least~$h$, as~$|Z| \leq \ell$. Let~$b'$ be a vertex of~$(X \cap \bigcap _{v \in A' \cup D} N_G(\phi_0(v))) \setminus Z$ that is not in the image of the partial model~$\phi'_1$ that is under construction. As~$\phi'_1$ does not yet specify an image for~$b$, its image contains at most~$h - 1$ vertices, which implies that such a vertex~$b'$ exists. Then define~$\phi'_1(b) := b'$. By the choice of~$b$ this image is contained in~$X \setminus Z$ and is adjacent to the images of all of~$b$'s neighbors in~$H$. \item After having defined~$\phi'_1$ on all vertices~$(A \cup D) \cap P_0$ and~$\hat{B}$ in the previous two steps, it remains to define~$\phi'_1$ on~$V(\hat{H}) \setminus P_0$. If~$V(\hat{H}) \setminus P_0 = \emptyset$ then the model is already completed and we are done. In the remaining case, the size of the current image of~$\phi'_1$ under construction is strictly less than~$h$. Observe that the set~$\Phi'$ contained exactly~$\ell + h$ partial $\hat{H}$-subgraph models such that for all distinct~$\phi, \phi' \in \Phi'$ we have~$\phi(V(\hat{H}) \setminus P_0) \cap \phi'(V(\hat{H}) \setminus P_0) = \emptyset$. Since~$Z$ has size at most~$\ell$, while the current image of~$\phi'_1$ has size less than~$h$, there is a full $\hat{H}$-subgraph model~$\phi^* \in \Phi$ such that~$\phi^*(V(\hat{H}) \setminus P_0)$ contains no vertex of~$Z$ and contains no vertex of the current image of~$\phi'_1$. We use~$\phi^*$ to define~$\phi'_1$ on the vertices in~$V(\hat{H}) \setminus P_0$: for each vertex~$v \in V(\hat{H}) \setminus P_0$ set~$\phi'_1(v) := \phi^*(v)$. Since~$X$ contains~$\phi(V(\hat{H}))$, all images assigned in this step are contained in~$X$. As~$\phi^*$ is a full $\hat{H}$-subgraph model that extends~$\phi_0$, its images realize all the edges of~$\hat{H}$. The only vertices of~$V(H) \setminus V(\hat{H})$ are those in~$\hat{B}$. As the vertices in~$\hat{B}$ have all their $H$-neighbors in~$P_0$, which~$\phi^*$ assigns the same images as~$\phi_0$, all edges of~$H$ are realized by the partial model~$\phi'_1$, which is contained in~$G[X] - Z$. \end{itemize} \noindent The construction of~$\phi'_1$ proves that~\ref{rep:set:subgr:preservesmodel} holds, which concludes the argumentation for Case 2.b. It is easy to see that the running time of the algorithm is polynomial in~$2^\alpha + |V(G)|^\mu$. There are two nontrivial computational steps: marking common neighbors for all subsets of~$A \cap P_0$ in Case 2.2 (which contributes the~$2^{\alpha}$ factor), and the invocation of Lemma~\ref{lemma:compute:extension:partial:separated:subgraph} for constructing the set~$\Phi$ in Case 2. As the time bound for Lemma~\ref{lemma:compute:extension:partial:separated:subgraph} is dominated by~$|V(G)|^{\mu}$, the algorithm obeys the claimed size bound. As the case distinction is exhaustive, this finishes the proof of Lemma~\ref{lemma:compute:representative:set:subgraph}. \end{proof} \subsection{Representative sets for small separators into small components} The proof of the following lemma can be considered as a straight-forward application of the sunflower lemma. We provide its proof for completeness, and to illustrate how it provides an analogue of Lemma~\ref{lemma:compute:representative:set:subgraph} for constant-size graphs~$H$. \begin{lemma} \label{lemma:compute:representative:set:constanth} There is an algorithm with the following specifications. The input is a graph~$G$, a graph~$H$ with a (possibly empty) vertex set~$D \subseteq V(H)$, an integer~$\ell$, and a partial subgraph model~$\phi_0$ of~$H$ with domain~$P_0 \supseteq D$. Let~$\alpha := |V(H) \setminus P_0|$ and let~$h := |V(H)|$. The output is a set~$X \subseteq V(G)$ with the following properties. \begin{enumerate}[(P1)] \item $|X| \leq h + (\alpha+1)! \cdot (\ell+1)^{\alpha}$.\label{rep:set:constanth:size} \item For any vertex set~$Z \subseteq V(G)$ of size at most~$\ell$, if~$G - Z$ contains a full subgraph model~$\phi_1$ of~$H$ that extends~$\phi_0$, then~$G[X] - Z$ contains a full subgraph model~$\phi'_1$ of~$H$ that extends~$\phi_0$.\label{rep:set:constanth:preservesmodel} \end{enumerate} The running time of the algorithm is~$|V(G)|^{\Oh(\alpha)}$. \end{lemma} \begin{proof} Define a system of sets~$\S$ as follows. For each full $H$-subgraph model~$\phi_1$ in~$G$ that extends~$\phi_0$, add a set~$S_{\phi_1} := \phi_1(V(H) \setminus P_0)$ to~$\S$ if it was not contained in~$\S$ already. The time bound allows us to try all images of~$V(H) \setminus P_0$ by brute force; we thus obtain the family~$\S$ in time~$|V(G)|^{\Oh(\alpha)}$. We will compute a set~$\S' \subseteq 2^{V(G)}$ with the following \emph{preservation property}: for every set~$Z \subseteq V(G)$ of size at most~$\ell$, if~$G - Z$ contains a full $H$-subgraph model~$\phi_1$ that extends~$\phi_0$, then~$G-Z$ contains a full $H$-subgraph model~$\phi'_1$ that extends~$\phi_0$ and satisfies~$\phi'_1(V(H) \setminus P_0) \in \S'$. By our choice of~$\S$ it is clear that~$\S$ has the preservation property, so we initialize~$\S'$ as a copy of~$\S$. \begin{claim} If~$\S' \subseteq \S$ has the preservation property and~$|\S'| \geq \alpha!(\ell+1)^{\alpha}$, we can identify a set~$S^* \in \S'$ in time~$|V(G)|^{\Oh(\alpha)}$ such that~$\S' \setminus \{S^*\}$ also has the preservation property. \end{claim} \begin{claimproof} Suppose that~$|\S'| \geq \alpha!(\ell+1)^{\alpha}$. By Lemma~\ref{lemma:sunflowers} there is a sunflower in~$\S'$ consisting of at least~$\ell+2$ sets~$S_1, \ldots, S_{\ell+1}, S_{\ell+2}$ and this can be found in time polynomial in the size of the set family and the universe. Let~$C := \bigcap _{i=1}^{\ell+2} S_i$ be the core of the sunflower. We show that~$\S' \setminus \{S_1\}$ has the preservation property. Let~$Z \subseteq V(G)$ have size at most~$\ell$ and suppose that~$\phi_1$ is a full $H$-subgraph model in~$G - Z$ that extends~$\phi_0$. As~$\S'$ has the preservation property, there is a full $H$-subgraph model~$\phi'_1$ in~$G-Z$ that extends~$\phi_0$ such that~$S_{\phi'_1} := \phi'_1(V(H) \setminus P_0)$ is contained in~$\S'$. If~$S_{\phi'_1} \neq S_1$ then the set~$S_{\phi'_1}$ is contained in~$\S' \setminus \{S_1\}$ which establishes the preservation property. If~$S_1 = S_{\phi'_1}$, then we proceed as follows. Since~$\phi'_1$ is an $H$-subgraph model in~$G - Z$, we have~$\phi'_1(V(H) \setminus P_0) \cap Z = S_1 \cap Z = \emptyset$ which shows in particular that none of the vertices in the core~$C$ of the sunflower are contained in~$Z$. Since the petals~$S_1 \setminus C, \ldots, S_{\ell + 2} \setminus C$ of the sunflower are pairwise disjoint, the set~$Z$ of size at most~$\ell$ can intersect at most~$\ell$ sets among~$S_2 \setminus C, \ldots, S_{\ell + 2} \setminus C$, implying that there is an~$i > 1$ such that~$(S_i \setminus C) \cap Z = \emptyset$ which implies~$S_i \cap Z = \emptyset$. Since~$i > 1$ this set~$S_i$ is contained in~$\S' \setminus \{S_1\}$. If~$\phi_1$ is a model in~$G-Z$ that extends~$\phi_0$ then~$\phi_0(P_0) \cap Z = \emptyset$. Let~$\phi_i$ be the $H$-subgraph model extending~$\phi_0$ that caused the set~$S_i$ to be added to~$\S$. Then~$\phi_i(V(H)) = \phi_i(V(H) \setminus P_0) \cup \phi_i(P_0) = S_i \cup \phi_i(P_0)$. As both sets are disjoint from~$Z$, we conclude that~$\phi_i$ is a full $H$-subgraph model in~$G-Z$ that extends~$\phi_0$ and satisfies~$\phi_i(V(H) \setminus P_0) \in \S' \setminus \{S_1\}$. Hence~$\S' \setminus \{S_1\}$ has the preservation property. \end{claimproof} By iterating the argument above, we arrive at a set system~$\S' \subseteq \S$ with the preservation property that contains at most~$\alpha!(\ell+1)^\alpha$ sets. The preservation property directly implies that picking~$X := \phi_0(P_0) \cup \bigcup _{S \in \S'} S$ satisfies~\ref{rep:set:constanth:preservesmodel}. Since each set in~$\S$ has size at most~$\alpha$ we find that~$|X| \leq |P_0| + \alpha \cdot \alpha!(\ell+1)^{\alpha}$, which satisfies~\ref{rep:set:constanth:size}. Since each iteration can be done in time polynomial in the size of~$\S$ and~$G$, while the number of iterations is bounded by~$|\S|$ which is~$\Oh(|V(G)|^\alpha)$, this concludes the proof. \end{proof} \subsection{General representative sets for subgraph detection} In this section, we put together Lemma~\ref{lemma:compute:representative:set:subgraph} and \ref{lemma:compute:representative:set:constanth} to obtain a marking procedure for $(a,b,c,d)$-splittable graphs. We show that if $D$ realizes the $(a,b,c,d)$-split and we invoke the marking procedure separately for each component of $H-D$, then the union of the marked sets contains a solution, if exists. \begin{lemma} \label{lemma:kernel:generic} For every triple of constants~$a,b,d \in \mathbb{N}$ there is an algorithm with the following specifications. The input is a graph~$G$, a graph~$H$ with a (possibly empty) vertex set~$D \subseteq V(H)$ that realizes an~$(a,b,|D|,d)$ split of~$H$, and a partial subgraph model~$\phi_0$ of~$H$ with domain~$P_0 = D$. Let~$k := |V(H)|$. The output is a set~$X \subseteq V(G)$ of size~$\Oh(k^{\Oh(a + b^2 + d)})$ such that~$G$ contains a full $H$-subgraph model that extends~$\phi_0$ if and only if~$G[X]$ contains a full $H$-subgraph model that extends~$\phi_0$. The running time of the algorithm is polynomial in~$|V(G)|$ for fixed~$a,b$, and~$d$. \end{lemma} \begin{proof} Fix~$a,b,d \in \mathbb{N}$ and let~$(G,H,D,\phi_0, P_0)$ be an input satisfying the requirements. Let~$C_1, \linebreak[1] \ldots, C_r$ be the connected components of~$H - D$ and define~$C'_i$ as the subgraph of~$H$ induced by~$N_H[V(C_i)]$ (i.e.,~$C'_i$ is component~$C_i$ together with its neighbors in~$D$). We build a set~$X$ as follows. Initialize~$X$ as~$\phi_0(D)$ and do the following for each~$i \in [r]$. \begin{itemize} \item If~$|V(C_i)| \leq a$ then we invoke the algorithm of Lemma~\ref{lemma:compute:representative:set:constanth} with parameters~$\hat{G},\hat{H},\hat{D},\hat{\ell},\hat{\phi}_0$ chosen as follows. The source graph~$\hat{G}$ equals~$G$, the query graph~$\hat{H}$ is set to~$C'_i$, the set~$\hat{D}$ is~$V(C'_i) \cap D$, the value of~$\hat{\ell}$ is~$k - |V(C'_i)|$, and~$\hat{\phi}_0$ is the restriction~$\phi_0|_{V(C'_i)}$. Under these definitions, the value of~$\hat{\alpha}$ defined in the lemma is~$|V(C'_i) \setminus (V(C'_i) \cap D)| = |V(C_i)| \leq a$. The algorithm outputs a set~$X_{C'_i}$ of size at most~$|V(C'_i)| + (a + 1)! \cdot (k - |V(C'_i)| + 1)^a$ which is~$\Oh(k^a)$ since~$a$ is a constant. We add~$X_{C'_i}$ to~$X^*$. \item If~$|V(C_i)| > a$, then by the precondition~$C_i$ is a $b$-thin bipartite graph. Let~$A$ and~$B$ the partite classes of~$H$ and choose~$A$ to be the smallest class, which has size at most~$b$. The precondition ensures that the number of vertices in~$C'_i$ that are not universal to~$V(C'_i) \cap D$ in the graph~$C'_i$ is at most~$d$. We invoke the algorithm of Lemma~\ref{lemma:compute:representative:set:subgraph} with parameters~$\hat{G} := G, \hat{H} := C'_i, \hat{D} := V(C'_i) \cap D, \hat{A} := A, \hat{B} := B, \hat{\ell} := k - |V(C'_i)|$, and we let~$\hat{\phi}_0$ be the restriction~$\phi_0|_{V(C'_i)}$. Therefore the value of~$\hat{\mu}$ defined in the lemma is bounded by~$|\hat{A}| + d + \sum _{v \in \hat{A}} |\hat{A}| \leq b + d + b^2$. By~\ref{rep:set:subgr:size} the size of the set~$X_{C'_i}$ computed by the described algorithm is bounded by~$(|V(C'_i)|^2 (k - |V(C'_i)|) + |V(C'_i)|^3)^{\hat{\mu} + 1} \cdot (1 + 2^b (k - |V(C'_i)| + |V(C'_i)|))$. Since~$b$ is a constant and~$|V(C_i)| \leq k$ we find that~$|X'_{C_i}| \in \Oh(k^{\Oh(\hat{\mu})}) \in \Oh(k^{\Oh(b^2 + d)})$. We add~$X'_{C_i}$ to~$X^*$. \end{itemize} Observe that the number of connected components of~$H$ does not exceed its order~$k$. Hence~$X^*$ is the union of~$\phi_0(D)$ with at most~$k$ sets that each have size~$\Oh(k^a + k^{\Oh(b^2 + d)})$, which proves that~$|X^*|$ is bounded by a fixed polynomial in the parameter~$k$. It is easy to verify that the running time of the procedure is polynomial in~$|V(G)|$ for fixed~$a,b$, and~$d$. To prove the correctness of the procedure, it remains to prove that~$G$ contains an $H$-subgraph model that extends~$\phi_0$ if and only if~$G[X]$ does. As~$G[X]$ is an induced subgraph of~$G$, the implication from right to left is trivial. We therefore consider the case that~$G[X]$ contains a full~$H$-subgraph model~$\phi^*$ that extends~$\phi_0$ and proceed to prove that~$G[X]$ contains such an extension as well. We construct an $H$-subgraph model~$\phi'$ in~$G[X]$ that extends~$\phi_0$ by repeatedly moving the $\phi^*$-model of one connected component of~$H - D$ into the set~$G[X]$. This is formalized by the following claim. \begin{claim} If~$\phi^*$ is a full $H$-subgraph model in~$G$ that extends~$\phi_0$, then for each~$i$ with~$0 \leq i \leq r$ the graph~$G$ contains a full $H$-subgraph model~$\phi_1$ that extends~$\phi_0$ and satisfies~$\phi_1(V(C'_j)) \subseteq X$ for all~$1 \leq j \leq i$. \end{claim} \begin{claimproof} We use induction on~$i$. For~$i=0$ the model~$\phi^*$ satisfies the stated condition. Consider some~$i > 0$ and assume by the induction hypothesis that~$\phi'_1$ is a full $H$-subgraph model that extends~$\phi_0$ and satisfies~$\phi'_1(V(C'_j)) \subseteq X$ for all~$1 \leq j \leq i - 1$. During the construction of~$X$ we considered the component~$C'_i$ and computed a set~$X_{C'_i}$ by invoking Lemma~\ref{lemma:compute:representative:set:constanth} or Lemma~\ref{lemma:compute:representative:set:subgraph}. Both lemmata guarantee that the set~$X_{C'_i}$ they output satisfies the following: if~$Z \subseteq V(G)$ has size at most~$\hat{\ell} = k - |V(C'_i)|$ and~$G - Z$ contains a full $C'_i$-subgraph model that extends~$\phi_0|_{V(C'_i)}$, then~$G[X_{C'_i}] - Z$ contains a full~$C'_i$-subgraph model~$\phi_{C'_i}$ that agrees with~$\phi_0|_{P_0 \cap V(C'_i)}$ on~$V(C'_i) \cap D$. Now choose~$Z := \phi'_1(V(H) \setminus V(C'_i))$, which has size~$k - |V(C'_i)|$, and observe that~$\phi'_1|_{V(C'_i)}$ is a full~$C'_i$ model in~$G - Z$ that extends~$\phi_0|_{V(C'_i)}$. Hence the left hand side of the implication mentioned above is satisfied and~$G[X_{C'_i}] - Z \subseteq G[X] - Z$ contains a full~$C'_i$-subgraph model~$\phi_{C'_i}$ that agrees with~$\phi_0|_{P_0 \cap V(C'_i)}$ on~$V(C'_i) \cap D$. Now build a model~$\phi_1$ as described in the claim, as follows. Initialize~$\phi_1$ as~$\phi'_1|_{V(H) \setminus V(C'_i)}$; by the induction hypothesis this maps the vertices of the first~$i - 1$ components into~$X$. Then augment the model~$\phi_1$ by mapping the vertices of~$V(C'_i)$ in the same way as~$\phi_{C'_i}$. Model~$\phi_{C'_i}$ maps all vertices into~$X - Z$, which implies that the vertices of~$C'_i$ are not mapped to vertices that are used in the model~$\phi'_1|_{V(H) \setminus V(C'_i)}$ as they are contained in~$Z$. Hence this extension results in an~$H$ subgraph model~$\phi_1$ in~$G[X]$ that maps the first~$i$ components into~$X$. It is a valid $H$-subgraph model since we are combining two partial models~$\phi_{C'_i}$ and~$\phi'_1|_{V(H) \setminus V(C_i)}$ for subgraphs~$C'_i$,~$V(H) \setminus V(C_i)$ of~$H$ that agree on the mapping of the separator~$V(C'_i) \cap D$, which may be verified using Observation~\ref{observation:merge:models}. To see that~$\phi_1$ extends~$\phi_0$, observe that~$\phi_1$ agrees with~$\phi_0$ on~$P_0 \setminus V(C'_i)$ as we copied the behavior on these vertices from~$\phi'_1$, which extends~$\phi_0$; it agrees with~$\phi_0$ on~$P_0 \cap V(C'_i)$ since we copied the behavior on these vertices from~$\phi_{C'_i}$ which agrees with~$\phi_0|_{P_0 \cap V(C'_i)}$ on~$V(C'_i) \cap D$. Hence it agrees with~$\phi_0$ on~$P_0$ and therefore extends~$\phi_0$. \end{claimproof} The case~$i = r$ of the claim proves that~$X$ preserves the existence of $H$-subgraph models extending~$\phi_0$. Note that a model as constructed in the claim for~$i = r$ maps all vertices of~$D$ into~$X$, as the model extends~$\phi_0$ while~$\phi_0(D) \subseteq X$ by construction. This concludes the proof of Lemma~\ref{lemma:kernel:generic}. \end{proof} \section{Kernelization complexity of packing problems} In this section, we prove Theorem~\ref{theorem:intro:packing} characterizing the hereditary classes $\F$ for which \FPacking admits a polynomial many-one or Turing kernel; the outcomes coincide for \FPacking. In Section~\ref{subsection:packing:upperbounds}, we invoke the marking algorithm developed in Section~\ref{sec:comp-repr-sets} to give a polynomial kernel for small/thin classes. In Section~\ref{subsection:packing:lowerbounds}, we establish a series of \textup{WK[1]}-hardness results for packing various basic classes of graphs. In Section~\ref{sec:packing-ramsey}, we prove Theorem~\ref{theorem:ramsey:1} characterizing small/thin hereditary classes. In Section~\ref{subsection:packing:theorem}, we put together all these results to complete the proof of Theorem~\ref{theorem:intro:packing}. \subsection{Upper bounds} \label{subsection:packing:upperbounds} To prove the positive part of Theorem~\ref{theorem:intro:packing} for small/thin hereditary classes, we obtain a kernel by invoking the marking algorithm of Lemma~\ref{lemma:kernel:generic} to obtain a bounded-size instance. \restatepackingkernel* \begin{proof} Let~$\F$ be a hereditary family for which such constants~$a$ and~$b$ exist. We show how to derive a polynomial many-one kernel for \kFPacking. An instance of \kFPacking consists of a tuple~$(G, H \in \F, t \in \mathbb{N})$ that asks whether~$G$ contains~$t$ vertex-disjoint subgraphs isomorphic to~$H$. Recall that the parameter is~$k := t \cdot |V(H)|$. Presented with an instance~$(G,H,t)$, the kernelization algorithm invokes Lemma~\ref{lemma:kernel:generic} to the graph~$G$, using~$t \cdot H$ as the query graph, an empty set~$D$, and an empty model~$\phi_0$. For fixed~$a$ and~$b$ the computation outputs a set~$X^* \subseteq V(G)$ in polynomial time. For the kernelization we output the instance~$(G[X^*], H, t)$. The lemma guarantees that~$|X^*| \in \Oh(k^{\Oh(a + b^2)})$ and that~$G[X^*]$ contains a $t \cdot H$-subgraph if and only if~$G$ does. Hence this procedure forms a correct kernelization. \end{proof} \subsection{Lower bounds} \label{subsection:packing:lowerbounds} In this section we present the polynomial-parameter transformations that establish Theorem~\ref{theorem:packing:lowerbounds}, which we repeat here for the reader's convenience. \restatepackinglower* The proofs are ordered in increasing difficulty. For each of the graph families mentioned in the theorem, we give a polynomial-parameter transformation from the \nRegularExactSetCover problem, whose \textup{WK[1]}-hardness we established in Lemma~\ref{lemma:regularexactsetcover:wkhard}. \begin{lemma} \label{lemma:fountainpacking:wkhard} \textsc{\Ffountain{s}-Packing} is \textup{WK[1]}-hard for any odd integer~$s \geq 3$. Similarly, \textsc{\Flongfountain{s}{t}-Packing} is \textup{WK[1]}-hard for any odd integer~$s \geq 3$ and any integer~$t \geq 1$. \end{lemma} \begin{proof} Let~$(r, \S, U, n)$ be an instance of \nRegularExactSetCover as described in Section~\ref{section:wkhardness}. We show how to construct an equivalent instance of a packing problem as described in the lemma statement in polynomial time, with a parameter that is polynomial in~$n$. An instance of the packing problem consists of a host graph~$G$, a pattern graph~$H$ that is contained in the relevant family~$\F$, and an integer~$t$. The construction we present works simultaneously for the family \Ffountain{s} for any odd~$s \geq 3$ and for the family \Flongfountain{s}{t} for any odd~$s \geq 3$ and~$t \geq 1$. To make the construction generic, let~$H$ be the (long) fountain in the class~$\F$ that we are targeting where the unique high-degree vertex of the (long) fountain has exactly~$r$ pendant vertices attached to it. The instance resulting from the transformation will ask for a packing of~$t := n / r$ copies of~$H$. Let~$c$ be the unique high-degree vertex in~$H$ to which the pendant vertices are attached. Let~$H'$ be the graph~$H$ without the~$n$ pendant vertices. The host graph~$G$ is defined as follows. \begin{itemize} \item Graph~$G$ contains the vertex set~$U$ as an independent set. Observe that we identify the elements of the universe~$U$ with a subset of the vertices of~$G$ by this definition. \item For every set~$S \subseteq U \in \S$, which has size exactly~$r$, we add a new copy~$H'_S$ of the graph~$H'$ to~$G$. Let~$c_S \in V(H'_S)$ be the copy of the vertex~$c$. We add edges between~$c_S$ and the vertices representing~$S$. This concludes the construction of~$G$. \end{itemize} It is easy to see that the construction can be carried out in polynomial time. The parameter~$k$ of the problem is defined as~$k := t \cdot |V(H)|$. Since~$|V(H)| \in \Oh(r)$ (the exact value depends on the constant~$s$ and possibly on~$t$) and~$t = n/r$ we have~$k \in \Oh(n)$, which is polynomially bounded in the parameter of the input instance. To complete the polynomial-parameter transformation it remains to prove that the instance~$(r, \S, U, n)$ of \nRegularExactSetCover is equivalent to the constructed instance~$(G,H,t)$ of \kFSubgraphTest. \begin{claim} If~$(\S,U)$ has an exact cover, then~$t \cdot H \subseteq G$. \end{claim} \begin{claimproof} Assume that~$(\S,U)$ has an exact cover, which must consist of~$t=n / r$ distinct and disjoint sets~$S_1, \ldots, S_t \in \S$. For each~$S_i$ with~$i \in [t]$ the copy~$H'_{S_i}$ of~$H'$ in~$G$ together with the~$r$ vertices~$S_i$ in~$G$ form a~$H$-subgraph in~$G$. Since the sets~$S_i$ are pairwise disjoint, there are~$t$ vertex-disjoint $H$-subgraphs in~$G$. Hence~$t \cdot H$ is a subgraph of~$G$. \end{claimproof} Before proving the reverse direction, we establish a structural claim. \begin{claim} \label{claim:fountain:hardness:goodmodels} Any~$H$ subgraph in~$G$ consists of a subgraph~$H'_S$ for some~$S \in \S$ together with the~$r$ vertices in~$N_G(c_S) \cap U$. \end{claim} \begin{claimproof} Recall that~$s \geq 3$ is odd. The only simple odd cycles in~$G$ are the copies of the unique length-$s$ cycle in~$H'$. To see this, pick an arbitrary edge~$e$ on the length-$s$ cycle in~$H'$ and let~$Y = \{e_1, \ldots, e_{|\S|}\}$ be the copies of this edge in the graphs~$H'_S$ for~$S \in \S$. Deleting the edges of~$Y$ from~$G$ results in a bipartite graph, which may be verified by noting that the graph~$G - Y$ can be reduced to the graph~$G[U \cup \{c_S \mid S \in \S\}]$ by repeatedly removing degree-one vertices (which does not remove any odd cycles), while~$G[U \cup \{c_S \mid S \in \S\}]$ is clearly bipartite as both~$G$ and~$\{c_S \mid S \in \S\}$ form independent sets in~$G$. So all odd cycles in~$G$ use an edge in~$Y$, and since each edge~$e_i \in Y$ is contained in a unique simple odd cycle (the cycle cannot be made longer without repeating a vertex, see Figure~\ref{fig:basic}) this shows that the only simple odd cycles in~$G$ are the length-$s$ cycles in the copies of~$H'$. Hence the only way an $H$-subgraph in~$G$ can realize a the length-$s$ cycle is to map it to a length-$s$ cycle in one of the copies~$H'$. Since the copies of~$H'$ are not connected to each other, while they only connect to~$U$ through their center vertex~$c_S$, an $H$-subgraph in~$G$ can only appear as described in the claim. \end{claimproof} \begin{claim} If~$t \cdot H \subseteq G$, then~$(\S,U)$ has an exact cover. \end{claim} \begin{claimproof} Assume that~$t \cdot H \subseteq G$. By Claim~\ref{claim:fountain:hardness:goodmodels}, every $H$-subgraph in~$G$ consists of a subgraph~$H'_S$ for~$S \in \S$ together with the~$r$ vertices in~$N_G(c_S) \cap U$. If there are~$t$ vertex-disjoint copies of~$H$ in~$G$, this implies that there are~$t$ vertices~$c_{S_1}, \ldots, c_{S_t}$ such that the sets~$N_G(c_{S_i})$ and~$N_G(c_{S_j})$ are disjoint for~$i \neq j$. As these sets contain~$t \cdot r = n$ vertices in total, the sets~$N_G(c_{S_1}) = S_1, \ldots, N_G(c_{S_t}) = S_t$ cover all of~$U$. The corresponding sets~$S_1, \ldots, S_t$ therefore form an exact cover of~$U$. \end{claimproof} As the claims establish the equivalence of the input and output instance, this concludes the proof of Lemma~\ref{lemma:fountainpacking:wkhard}. \end{proof} \begin{lemma} \label{lemma:operahousepacking:wkhard} \textsc{\Foperahouse{s}-Packing} is \textup{WK[1]}-hard for any odd integer~$s \geq 1$. \end{lemma} \begin{proof} Let~$s\geq 1$ be an odd integer. We transform an instance~$(r, \S, U, n)$ of \nRegularExactSetCover in polynomial time into an equivalent instance~$(G, H := \Goperahouse{s}{r},t := n/r)$ of \textsc{\Foperahouse{s}-Packing} with~$k := t \cdot |V(H)| \in \Oh(n)$ (see Figure~\ref{fig:basic}). The host graph~$G$ is defined as follows. \begin{itemize} \item Graph~$G$ contains the vertex set~$U$ as an independent set. \item For every size-$r$ set~$S \in \S$ we add a new length-$s$ path~$P_S$ to~$G$. Let~$x_S$ and~$y_S$ be the endpoints of~$P_S$. We make the vertices~$x_S$ and~$y_S$ adjacent to all vertices representing members of~$S$. \end{itemize} As the construction can be performed in polynomial time and produces an instance of the target problem whose parameter is suitably bounded, it remains to prove that the input and output instances are equivalent. \begin{claim} If~$(\S,U)$ has an exact cover, then~$t \cdot H \subseteq G$. \end{claim} \begin{claimproof} Assume that~$(\S,U)$ has an exact cover~$S_1, \ldots, S_t \in \S$. For each~$S_i$ with~$i \in [t]$ the path~$P_{S_i}$ together with the~$r$ vertices~$S_i$ form a~$H$-subgraph in~$G$. Since the sets~$S_i$ are pairwise disjoint, there are~$t$ vertex-disjoint $H$-subgraphs in~$G$. Hence~$t \cdot H$ is a subgraph of~$G$. \end{claimproof} \begin{claim} \label{claim:operahouse:hardness:goodmodels} Any~$H$ subgraph in~$G$ consists of a path~$P_S$ for some~$S \in \S$ together with the~$r$ vertices in~$N_G(x_S) \cap N_G(y_S) = S$. \end{claim} \begin{claimproof} Recall that~$s \geq 1$ is odd. The only simple odd cycles in~$G$ of length~$s+2$ are those consisting of a path~$P_S$ for~$S \in \S$ together with a vertex in~$N_G(x_S) \cap N_G(y_S) \subseteq U$. To see this, observe that removing the first edge of every path~$P_S$ ($S \in \S$) results in a bipartite graph: after removing these edges and iterately removing degree-one vertices, we are left with the bipartite graph where~$U$ forms one partite class and~$\{x_S, y_S \mid S \in \S\}$ forms the other partite class. Hence every odd cycle contains an edge on a path~$P_S$, and use of such an edge forces the entire path~$P_S$ to be used. As it has length~$s$, to complete a length-$(s+2)$ path requires a vertex in the common neighborhood of the endpoints of the path; as~$s$ is odd, the only such vertices are those in~$N_G(x_S) \cap N_G(y_S) = U$. Now consider a $H$-subgraph in~$G$. As~$H = \Goperahouse{s}{r}$ contains a length~$(s+2)$ cycle, the above argument shows that the model of this cycle is formed by a path~$P_S$ for~$S \in \S$ together with a common $U$-neighbor of~$\{x_S, y_S\}$, resulting in a length-$(s+2)$ cycle~$C'$ in~$G$. Since the vertices~$x_S$ and~$y_S$ form the unique pair of vertices on~$C'$ that have degree at least~$r + 1$ in~$G$ and are connected by a path of length~$s$ through~$C'$, the two high-degree vertices of~$\Goperahouse{s}{r}$ must be realized by~$x_S$ and~$y_S$. Since~$N_G(x_S) \cap N_G(y_S) = S \subseteq U$, the subgraph must use the~$r$ vertices in~$S$ to realize the common neighbors of the endpoint of the path. \end{claimproof} \begin{claim} If~$t \cdot H \subseteq G$, then~$(\S,U)$ has an exact cover. \end{claim} \begin{claimproof} Assume that~$t \cdot H \subseteq G$. By Claim~\ref{claim:operahouse:hardness:goodmodels}, every $H$-subgraph in~$G$ consists of a path~$P_S$ for~$S \in \S$ together with the vertices representing~$S$. If there are~$t$ vertex-disjoint copies of~$H$ in~$G$, then there are~$t$ disjoint sets~$S_1, \ldots, S_r \in \S$. As these sets contain~$t \cdot r = n$ vertices in total, they cover all of~$U$ and form an exact cover. \end{claimproof} This concludes the proof of Lemma~\ref{lemma:operahousepacking:wkhard}. \end{proof} \begin{lemma} \label{lemma:substarpacking:wkhard} \textsc{\Fsubdivstar-Packing} is \textup{WK[1]}-hard. \end{lemma} \begin{proof} We transform an instance~$(r, \S, U, n)$ of \nRegularExactSetCover in polynomial time into an equivalent instance~$(G, H := \Gsubdivstar{r},t := n/r)$ of \textsc{\Fsubdivstar-Packing} with~$k := t \cdot |V(H)| \in \Oh(n)$. By the definition of \nRegularExactSetCover we have~$r \geq 3$. Let~$c \in V(H)$ be the center of the star whose subdivision yields~$H$, implying that~$\deg_H(c) \geq 3$. The host graph~$G$ is defined as follows. \begin{itemize} \item Graph~$G$ contains the vertex set~$U$ as an independent set. For every vertex~$u \in U$ we add a pendant degree-one neighbor~$u'$ adjacent to~$u$. \item For every size-$r$ set~$S \in \S$ we add a new vertex~$x_S$ to~$G$ that is adjacent to the~$r$ vertices in~$U$ representing~$S$. \end{itemize} As the construction can be performed in polynomial time and produces an instance of the target problem whose parameter is suitably bounded, it remains to prove that the input and output instances are equivalent. \begin{claim} If~$(\S,U)$ has an exact cover, then~$t \cdot H \subseteq G$. \end{claim} \begin{claimproof} Assume that~$(\S,U)$ has an exact cover~$S_1, \ldots, S_t \in \S$. For each~$S_i$ with~$i \in [t]$ the vertices~$\{u, u' \mid u \in S\}$ form a subdivided star with~$r$ leaves and~$x_{S_i}$ as its center. Since the sets~$S_i$ are pairwise disjoint, there are~$t$ vertex-disjoint $H$-subgraphs in~$G$. \end{claimproof} \begin{claim} \label{claim:subdivstar:hardness:goodmodels} If~$\phi$ is a full $H$-subgraph model in~$G$, then the following holds. \begin{itemize} \item If~$\phi(c) \in \{x_S \mid S \in \S\}$ then~$|\phi(V(H)) \cap U| \geq r$. \item If~$\phi(c) \not \in \{x_S \mid S \in \S\}$ then~$|\phi(V(H)) \cap U| \geq r + 1$. \end{itemize} \end{claim} \begin{claimproof} Consider a full $H$-subgraph model~$\phi$ in~$G$. Since~$\deg_H(c) \geq 3$ while~$\deg_G(u') = 1$ for all~$u \in U$, the image of~$c$ can only be a vertex in~$U$ or a vertex~$x_S$ for~$S \in \S$. If~$\phi(c) = x_S$ for some~$S \in \S$, then~$N_G(x_S) = S \subseteq \phi(V(H))$, as~$c$ has~$r$ neighbors in~$H$ that must be realized by the unique set of~$r$ neighbors of~$x_S$ in~$G$. This proves the first part of the claim. To prove the second part, by the observation above it suffices to consider the case that~$\phi(c) = u \in U$. Observe that the vertex~$u'$ has degree one in~$G$ and therefore cannot form the image of a subdivider vertex of the star. Hence all~$r$ subdivider vertices are realized by other neighbors of~$u$, which all belong to~$\{x_S \mid S \in \S\}$. Consequently, the images of the degree-one endpoints of the subdivided star are neighbors of~$\{x_S \mid S \in \S\}$ distinct from~$u$. As all neighbors of~$\{x_S \mid S \in \S\}$ belong to~$U$, this implies that~$|\phi(V(H)) \cap U| \geq r+1$. \end{claimproof} \begin{claim} If~$t \cdot H \subseteq G$, then~$(\S,U)$ has an exact cover. \end{claim} \begin{claimproof} Assume that~$t \cdot H \subseteq G$. By Claim~\ref{claim:subdivstar:hardness:goodmodels}, every $H$-subgraph model in~$G$ uses at least~$r$ vertices of~$U$. As there are only~$n$ vertices in~$U$ in total, to realize~$t = n/r$ vertex-disjoint $H$-subgraphs, each of the~$t$ subgraphs has to use exactly~$r$ vertices of~$U$. Such subgraphs therefore map the center of the subdivided star to a vertex~$x_S$ for~$S \in \S$. Hence there are~$t$ distinct vertices~$x_{S_1}, \ldots, x_{S_t}$ forming the centers of $H$-subgraph models in~$G$. For each~$i \in [t]$, the vertices~$N_G(x_{S_i}) = S_i$ must be used in the model containing~$x_{S_i}$, since they are the unique~$r$ neighbors of~$x_{S_i}$ in~$G$. Thus~$S_1, \ldots, S_t$ are pairwise disjoint sets in~$\S$ covering~$n$ elements in total, which means they form an exact cover of~$U$. \end{claimproof} This concludes the proof of Lemma~\ref{lemma:substarpacking:wkhard}. \end{proof} \begin{lemma} \label{lemma:broompacking:wkhard} \textsc{\Fdoublebroom{s}-Packing} is \textup{WK[1]}-hard for any odd integer~$s \geq 1$. \end{lemma} \begin{proof} We transform an instance~$(r \geq 3, \S, U, n)$ of \nRegularExactSetCover in polynomial time into an equivalent instance~$(G, H := \Gdoublebroom{s}{r},t := n/r)$ of \textsc{\Fdoublebroom{s}-Packing} with~$k := t \cdot |V(H)| \in \Oh(n)$. The host graph~$G$ is defined as follows. \begin{itemize} \item Graph~$G$ contains the vertex set~$U$ as an independent set. \item For every size-$r$ set~$S \in \S$ we add a new length-$s$ path~$P_S$ to~$G$. Let~$x_S$ and~$y_S$ be the endpoints of~$P_S$. We make~$x_S$ adjacent to the~$r$ vertices in~$U$ representing~$S$. We add~$r$ new vertices and make them adjacent to~$y_S$. \end{itemize} As the construction can be performed in polynomial time and produces an instance of the target problem whose parameter is suitably bounded, it remains to prove that the input and output instances are equivalent. \begin{claim} If~$(\S,U)$ has an exact cover, then~$t \cdot H \subseteq G$. \end{claim} \begin{claimproof} Assume that~$(\S,U)$ has an exact cover~$S_1, \ldots, S_t \in \S$. For each~$S_i$ with~$i \in [t]$ the vertices representing~$S$ combine with the path~$P_S$ and the~$r$ degree-one neighbors of~$y_S$ to form a \Gdoublebroom{s}{r} subgraph. Since the sets~$S_i$ are pairwise disjoint, there are~$t$ vertex-disjoint $H$-subgraphs in~$G$. \end{claimproof} Let~$x$ and~$y$ be the unique two vertices of degree~$r + 1$ in \Gdoublebroom{s}{r} (recall that~$s \geq 1$ and~$r \geq 3$). \begin{claim} \label{claim:doublebroom:hardness:goodmodels} If~$\phi$ is a full $H$-subgraph model in~$G$, then~$\phi(V(H)) \cap U \supseteq S$ for some~$S \in \S$. \end{claim} \begin{claimproof} Consider a full $H$-subgraph model~$\phi$ in~$G$. If~$\phi(\{x,y\}) = \{x_S, y_S\}$ for~$S \in \S$, then as~$P_S$ is the unique length-$s$ path in~$G$ between~$x_S$ and~$y_S$, this path is used as the image of the length-$s$ path in~$H$. Hence the unique~$r$ vertices~$N_G(x_S) \setminus V(P_S) = S$ must be used as images of the degree-one neighbors of~$x$ in~$H$. Hence~$\phi(V(H)) \supseteq S$. If~$\phi(\{x,y\}) \cap \{y_S \mid S \in \S\} \neq \emptyset$, then we must have~$\phi(\{x,y\}) = \{x_S, y_S\}$ for some~$S \in \S$, since the only vertex in~$G$ at distance exactly~$s$ from~$y_S$ that has degree more than one is~$x_S$. Hence if $\phi(\{x,y\}) \cap \{y_S \mid S \in \S\} \neq \emptyset$ then the previous argument shows that the claim holds. It remains to consider the case that the images of~$x$ and~$y$ do not belong to~$\{y_S \mid S \in \S\}$. As~$\deg_G(\phi(x)) \geq \deg_H(x) \geq 4$ (recall~$r \geq 3$) and~$\deg_G(\phi(y)) \geq 4$, the only remaining options for~$\phi(x)$ and~$\phi(y)$ are the vertices in~$U \cup \{x_S \mid S \in \S\}$. As the constructed graph~$G$ is bipartite and all vertices of~$U$ are contained in the same partite class, while the path connecting~$x$ and~$y$ in~$H$ has odd length~$s$, it follows that the model of one of~$\{x,y\}$ belongs to~$U$ while the other does not. By the previous argument, this implies that the model of the other vertex belongs to~$\{x_S \mid S \in \S\}$. Assume without loss of generality that~$\phi(y) \in U$ and~$\phi(x) = x_S$ for some~$S \in \S$. Since~$\deg_H(x) = \deg_G(x_S) = r + 1$, all vertices of~$N_G(x_S)$ are used in the model to realize the images of~$N_H(x)$. Since~$S \subseteq N_G(x_S)$, the claim follows. \end{claimproof} \begin{claim} If~$t \cdot H \subseteq G$, then~$(\S,U)$ has an exact cover. \end{claim} \begin{claimproof} Assume that~$t \cdot H \subseteq G$. By Claim~\ref{claim:doublebroom:hardness:goodmodels}, the image of every $H$-subgraph model in~$G$ is a superset of some~$S \in \S$. If there is a subgraph model that uses a strict superset, then this leaves less than~$n - r$ vertices to realize the remaining~$t-1$ $H$-subgraph models, which is not possible since each requires a superset of a size-$r$ set. Hence the intersection of each $H$-subgraph in the packing with~$U$ is a set~$S \in \S$. Since the packing consists of vertex-disjoint subgraphs, there are~$t$ pairwise disjoint sets in~$\S$. As they each have size~$r$, these~$n/r$ sets cover the entire universe~$U$. Hence~$(\S,U)$ has an exact cover. \end{claimproof} This concludes the proof of Lemma~\ref{lemma:broompacking:wkhard}. \end{proof} As we have given \textup{WK[1]}-hardness proofs for all graph families listed in Theorem~\ref{theorem:packing:lowerbounds}, the theorem follows. \subsection{Combinatorial characterizations} \label{sec:packing-ramsey} In this section, our goal is to prove Theorem~\ref{theorem:ramsey:1} characterizing hereditary classes that are not small/thin. If a class $\F$ is not small/thin, then $\F$ contains graphs either with large nonbipartite components or with large non-thin bipartite components. We investigate both cases separately and show that one the basic classes listed in Theorem~\ref{theorem:ramsey:1} is contained in $\F$. Let us treat first the case of large nonbipartite graphs. \begin{lemma} \label{lemma:ramsey:nonbipartite0} For every $Q\ge 1$, there is an $a=a(Q)\ge 1$ such that if $H$ is a connected nonbipartite graph with at least $a$ vertices, then $H$ contains at least one of the following graphs as induced subgraph: \begin{enumerate} \item $\Gpath{Q}$, \item $\Gclique{Q}$, \item $\Gfountain{s}{Q}$ for some odd integer $3\le s\le Q+2$, \item $\Glongfountain{s}{t}{Q}$ some odd integer $3\le s\le Q+2$ and integer $1\le t\le Q$, or \item $\Goperahouse{s}{Q}$ for some odd integer $1\le s\le Q$. \end{enumerate} \end{lemma} \begin{proof} We define the following constants: \begin{align*} d_1=&R(3Q,2)+(Q+1),\\ d_2=&R(3Q,2)+1+(Q+2)(d_1+1),\\ a=&\sum_{i=0}^{Q-1}d_2^i+1, \end{align*} where $R(3Q,2)$ is the Ramsey number guaranteeing a clique or independent set of size at least $3Q$ (Theorem~\ref{theorem:ramsey}). Observe that the constant $a$ depends only on $Q$. Let $C$ be an odd cycle of minimum length in $H$. Observe that $C$ has no chord, otherwise the chord would split the odd cycle $C$ into a strictly shorter odd cycle and a strictly shorter even cycle, a contradiction. Thus $C$ is an induced cycle and hence has length at most $Q+2$, otherwise $\Gpath{Q}$ would be an induced subgraph of $H$. Assume that $H$ has at least $a$ vertices, but does not contain any of the graphs listed in the lemma as induced subgraph. Select an arbitrary vertex of $H$. If $H$ does not contain $\Gpath{Q}$ as an induced subgraph, then every vertex of $H$ is at distance at most $Q-1$ from the selected vertex. If $H$ has maximum degree $d_2$, then this would imply that $H$ has at most $\sum_{i=0}^{Q-1}d_2^i<a$ vertices, a contradiction. Thus we may assume that $H$ has a vertex $v$ of degree at least $d_2$. We consider two cases. \textbf{Case 1: Every vertex on the cycle $C$ has degree at most $d_1$.} As $d_1<d_2$, vertex $v$ is not on $C$. Let $p_0=v$, $p_1$, $\dots$, $p_{\ell}=w$ be a shortest path $P$ from $v$ to a vertex $w$ of the cycle $C$. Vertex $v$ has at most $|C|(d_1+1)$ neighbors in the closed neighborhood of $C$ and (as $P$ is a shortest path), only one neighbor on $P$, namely $p_1$. Thus we can select from the at least $d_2$ neighbors of $v$ a set $X_1$ of at least $d_2-1-|C|(d_1+1)\ge d_2-1-(Q+2)(d_1+1)=R(3Q,2)$ vertices that are not on $P$, not on $C$, and not in the neighborhood of $C$. As $H$ has no clique of size $3Q$, Ramsey's Theorem (Theorem~\ref{theorem:ramsey}) implies that there is an independent set $X_2\subseteq X_1$ of size at least $3Q$. As $P$ is a shortest path, a vertex of $X_2$ cannot be adjacent to $p_i$ for $i>2$. Let us partition $X_2$ into three classes: let $X_{2,i}$ contain a vertex of $X_2$ if it is adjacent to $p_i$, but not to $p_{i'}$ for any $i'>i$ (as $p_0=v$, every vertex of $X_2$ is in exactly one of $X_{2,0}$, $X_{2,1}$, or $X_{2,2}$). As $|X_2|\ge 3Q$, there is an $0\le i^* \le 2$ such that $|X_{2,i^*}|\ge Q$. Note that $i^*<\ell$: since $p_\ell$ is on $C$, no vertex of $X_2$ is adjacent to it. As $P$ is a shortest path, only $p_{\ell-1}$ can have neighbors on the cycle $C$. We consider two subcases (see Figure~\ref{fig:oddcycle}). \begin{figure}[t] \begin{center} {\small \svg{\linewidth}{oddcycle}} \caption{Proof of Lemma~\ref{lemma:ramsey:nonbipartite0}.}\label{fig:oddcycle} \end{center} \end{figure} \textbf{Case 1.a: $p_\ell$ is the only neighbor of $p_{\ell-1}$ on $C$.} Then the set $X_{2,i^*}$, the subpath $P'$ of $P$ from $p_{i^*}$ to $p_\ell=w$, and the cycle $C$ form a $\Glongfountain{|C|}{|P'|}{|X_{2,i^*}|}$, contradicting our assumptions (note that $|C|\le Q+2$, $|P'|\le Q$, and $|X_{2,i^*}|\ge Q$). \textbf{Case 1.b: $p_\ell$ has at least two neighbors on $C$.} Then $C$ has a subpath $R$ of odd length such that $p_{\ell-1}$ is adjacent to the endpoints of $R$, but not to the internal vertices of $R$ (this is because the neighbors of $p_{\ell-1}$ split the cycle $C$ into subpaths, and at least one such subpath has odd length). We may observe that, by the minimality of the cycle $C$, the length of $R$ is exactly $|C|-2$, otherwise $R$ and $p_{\ell-1}$ would form an odd cycle strictly shorter than $C$. Now $X_{2,i^*}$, the subpath $P'$ of $P$ from $p_{i^*}$ to $p_{\ell-1}$, and path $R$ form a $\Glongfountain{|R|+2}{|P'|}{|X_{2,i^*}|}$. As $|R|+2= |C|\le Q+2$, $|P'|\le Q$, and $|X_{2,i^*}|\ge Q$, this contradicts assumptions on $Q$. \textbf{Case 2: $C$ has a vertex $w$ of degree at least $d_1$.} Vertex $w$ has a set $X_1$ of at least $d_1-(|C|-1)\ge d_1-(Q+2-1)=R(3Q,2)$ neighbors that are not on $C$. As $H$ has no clique of size $3Q$, Ramsey's Theorem (Theorem~\ref{theorem:ramsey}) implies that there is an independent set $X_2\subseteq X_1$ of size at least $3Q$. We need to treat the case when $C$ has length 3 separately, thus we consider two subcases (see Figure~\ref{fig:oddcycle}). \textbf{Case 2.a: $|C|=3$.} Let $w$, $z_1$, and $z_2$ be the vertices of $C$. If $X_2$ has a subset $X_3$ of $Q$ vertices adjacent to $z_1$, then the set $X_3$ and the path $wz_1$ form an $\Goperahouse{1}{Q}$. We can get an $\Goperahouse{1}{|Q|}$ in a similar way if $X_2$ has a subset $X_3$ of $Q$ vertices adjacent to $z_2$. Otherwise, there is a set $X_3\subseteq X_2$ of at least $Q$ vertices that are adjacent to neither $z_1$ nor $z_2$, and hence $X_3$ and the cycle $C$ form a $\Gfountain{3}{Q}$. \textbf{Case 2.b: $|C|\ge 5$.} Let $q_1$, $q_2$, $w$, $q_3$, $q_4$ be consecutive vertices appearing on $C$ in this order; as $|C|\ge 5$, these 5 vertices are distinct. It is easy to see that no vertex $u$ of $X_2$ can be adjacent to any vertex $q'$ of $C\setminus \{q_1,q_2,w,q_3,q_4\}$: then the path $wuq'$ and the subpath of $C$ between $w$ and $q'$ that has odd length would form an odd cycle strictly shorter than $C$, a contradiction. If $X_2$ has a vertex adjacent to either $q_2$ or $q_3$, then there is a triangle, contradicting the minimality of $C$. If $X_2$ has a vertex $u$ adjacent to both $q_1$ and $q_4$, then the path $q_1uq_4$ and the subpath of $C$ between $q_1$ and $q_4$ avoiding $w$ would form an odd cycle strictly shorter than $C$. Thus a vertex in $X_2$ can have at most one neighbor on $C$ besides $w$, and this neighbor can only be $q_1$ or $q_4$. If there is a set $X_3\subseteq X_2$ of $Q$ vertices that are adjacent to $q_1$, then the subpath of $C$ between $q_1$ and $w$ avoiding $q_2$ form an $\Goperahouse{|C|-2}{Q}$ (note that $|C|-2\le Q$). We get an $\Goperahouse{|C|-2}{Q}$ in a similar way if there is a set $X_3\subseteq X_2$ of $Q$ vertices that are adjacent to $q_4$. Otherwise, there is a set $X_3\subseteq X_2$ of $Q$ vertices that are adjacent to only $w$ on $C$; then $X_3$ and $C$ form a $\Gfountain{|C|}{Q}$. \end{proof} Lemma~\ref{lemma:ramsey:nonbipartite0} is a statement about a specific graph. We turn it into a statement on graph classes. \begin{lemma} \label{lemma:ramsey:nonbipartite} Let~$\F$ be a hereditary graph family. Then at least one of the following is true. \begin{enumerate} \item There is an integer $a\ge 1$ such that every connected nonbipartite graph in $\F$ has at most~$a$ vertices. \item $\F$ is a superset of \Fpath. \item $\F$ is a superset of \Fclique. \item $\F$ is a superset of \Ffountain{s} for some odd integer $s\ge 3$. \item $\F$ is a superset of \Flongfountain{s,t} for some odd integer $s\ge 3$ and integer $t\ge 1$. \item $\F$ is a superset of \Foperahouse{s} for some odd integer $s\ge 1$. \end{enumerate} \end{lemma} \begin{proof} Assuming that $\F$ is not a superset of $\Fpath$, there is an integer $\ell\ge 1$ such that $\Gpath{\ell}\not\in \F$. Then, assuming that $\F$ is not a superset of any of the other families described in the lemma, there is a $Q\ge \ell \ge 1$ such that $\F$ does not contain any of the following graphs: \begin{itemize} \item $\Gclique{Q}$, \item $\Gfountain{s}{Q}$ for any odd integer $3\le s < \ell+2$, \item $\Glongfountain{s}{t}{Q}$ for any odd integer $3 \le s < \ell+2$ and integer $1\le t < \ell$, \item $\Goperahouse{s}{Q}$ for any odd integer $1\le s < \ell$. \end{itemize} The reason why such a $Q$ can be defined is that we have to consider a finite number of infinite sequences of graphs (such as $\Gfountain{s}{i}$ for a fixed $3\le s \le \ell+2$ and for $i=1,2,\dots$), we know by assumption that these sequences are not contained in $\F$, hence for each sequence there is an $i$ such that the $i$-th element of the sequence is not contained in $\F$. In fact, as the $i$-th element of the sequence is an induced subgraph of the $i'$-th element for every $i'>i$, we know that no element of the sequence after the $i$-th element can appear in $\F$ either. Then we can define $Q$ to be the maximum of all these $i$'s corresponding to the finitely many forbidden sequences. As $\F$ is hereditary, we also know that none of these graphs appear as induced subgraphs in the members of $\F$. Observe that an induced $\Gpath{\ell}$ is contained in every $\Gfountain{s}{Q}$ with $s\ge \ell+2$, in every $\Glongfountain{s}{t}{Q}$ with $s\ge \ell+2$ or $t\ge \ell$, and in every $\Goperahouse{s}{Q}$ with $s\ge \ell$. Therefore, none of the graphs listed in Lemma~\ref{lemma:ramsey:nonbipartite0} can appear in $\F$, hence every connected nonbipartite graph in $\F$ has size at most $a(Q)$. \end{proof} Next we consider large bipartite graphs that are not thin. \begin{lemma} \label{lemma:ramsey:balancedbipartite0} For every $Q\ge 1$, there is an integer $b=b(Q)\ge 1$ such that every connected bipartite graph is either $b$-thin or contains one the following graphs as induced subgraphs: \begin{enumerate} \item $\Gpath{Q}$, \item $\Gclique{Q}$, \item $\Gbiclique{Q}$, \item $\Gsubdivstar{Q}$, or \item $\Gdoublebroom{s}{Q}$ some odd integer $1\le s\le Q$. \end{enumerate} \end{lemma} \begin{proof} We define the following constants: \begin{align*} h&=P(Q,Q), && \textup{($P$ is from Corollary~\ref{theorem:path:ramsey2})}\\ \kappa&=\sum_{i=0}^{h}(Q-1)^i,\\ d&=1+R(\bipRamsey(2Q,2),2),&&\textup{($R$ is from Ramsey's Theorem; $\bipRamsey$ is from Theorem~\ref{theorem:bipartite:ramsey})}\\ b&=\kappa d. \end{align*} Observe that the constant $b$ depends only on $Q$. Assume that $H\in \F$ is a connected $b$-thin bipartite graph not containing any of the listed graphs as an induced subgraph. First we bound the vertex cover number of $H$. \begin{claim}\label{claim:vcbound} $H$ has a vertex cover of size at most $\kappa$. \end{claim} \begin{proof} Construct a DFS tree of $H$ starting at an arbitrary root. Let $L\subseteq V(H)$ be the set of leaves in the DFS tree and let $Z=V(H)\setminus L$. Observe that $Z$ is a vertex cover of $H$: there are no edges between the leaves. We claim that if a vertex $v\in Z$ has $Q$ children $x_1,\dots,x_Q\in Z$ in the DFS tree, then $H$ contains $\Gsubdivstar{Q}$ as an induced subgraph, a contradiction. To show this, let us select an arbitrary child $y_i$ of each $x_i$ (note that $x_i\in Z$ is not a leaf). Then it is clear that the set $\{v,x_1,\dots, x_s,y_1,\dots,y_s\}$ forms a $\Gsubdivstar{Q}$ subgraph. To see that it is an induced subgraph, observe that there is no edge between $\{x_i,y_i\}$ and $\{x_j,y_j\}$ for any $i\neq j$ by the properties of the DFS tree and there is no edge $\{v,y_i\}$, as this would create a triangle $\{v,x_i,y_i\}$ and the graph is bipartite by assumption. Therefore, every $v\in Z$ has at most $Q-1$ children in $Z$, otherwise the graph would contain $\Gsubdivstar{Q}$ as an induced subgraph. The height of the DFS tree is at most $h=P(Q,Q)$, otherwise Corollary~\ref{theorem:path:ramsey2} would imply that $H$ contains $\Gpath{Q}$, $\Gclique{Q}$, $\Gbiclique{Q}$ as an induced subgraph. This means that $|Z|\le \sum_{i=0}^{h}(Q-1)^i=\kappa$. As we have observed, $Z$ is a vertex cover of $H$, thus $H$ has vertex cover number at most $\kappa$. \cqed \end{proof} Assume first that one partite class of $H$ has maximum degree $d$. Let $A$ and $B$ be the two partite classes and assume that the vertices in $A$ have degree at most $d$. As $H$ is not $b$-thin, both partite classes of $H$ have size more than $b$. Pick one edge incident to each vertex of $B$, we get $|B|\ge b+1=\kappa d+1$ distinct edges. Now each vertex can cover at most $d$ of these edges: a vertex in $A$ can cover at most $d$ (as it has degree at most $d$) and a vertex in $B$ covers exactly one such edge. This contradicts Claim~\ref{claim:vcbound}. Therefore, we can assume that $H$ has two vertices $x$ and $y$ having degree at least $d$ in each partite class of $H$. Let $p_0=x$, $p_1$, $\dots$, $p_{\ell-1}$, $p_\ell=y$ be a shortest path $P$ between $x$ and $y$ (see Figure~\ref{fig:2broom}). \begin{figure}[t] \begin{center} {\small \svg{0.7\linewidth}{2broom}} \end{center} \caption{Proof of Lemma~\ref{lemma:ramsey:balancedbipartite0}: the case when $|X_4|\ge Q$ and $|X_3\setminus X_4|\ge Q$.}\label{fig:2broom} \end{figure} As $x$ and $y$ are in different partite classes, we have that $\ell\ge 1$ is odd and $|P|<Q$, otherwise it would form an induced $\Gpath{Q}$. Let $X_1$ be a set of $d-1$ neighbors of $x$ different from $p_1$ and let $Y_1$ be a set of $d-1$ neighbors of $y$ different from $p_{\ell-1}$. Note that $X_1$ and $Y_1$ are in different partite classes, hence disjoint. Moreover, no vertex of $X_1$ or $Y_1$ is on the path $P$, as $P$ is a shortest path between $x$ and $y$. The definition of $d$ and Ramsey's Theorem (Theorem~\ref{theorem:ramsey}) implies that $X_1$ has a clique or independent set of size $\bipRamsey(2Q,2)$. As $H$ does not contain $\Gclique{Q}$ and certainly $Q$ is less than $\bipRamsey(2Q,2)$, the only possibility is that there is an independent set $X_2\subseteq X_1$ of size $\bipRamsey(2Q,2)$. Similarly, there is an independent set $Y_2\subseteq Y_1$ of size $\bipRamsey(2Q,2)$. As $H$ contains no $\Gbiclique{2Q}$ as induced subgraph, Theorem~\ref{theorem:bipartite:ramsey} implies that there are sets $X_3\subseteq X_2$ and $Y_3\subseteq Y_2$ of size $2Q$ such that there is no edge between $X_3$ and $Y_3$. It is clear that the sets $X_3$, $Y_3$, and the path $P$ form a $\Gdoublebroom{|P|}{Q}$ subgraph, but it is not necessarily an induced subgraph. As $P$ is a shortest $x-y$ path, there is no edge between $X_3$ and $p_i$ for $i\ge 3$, and there is no edge between $X_3$ and $p_1$, as it would form a triangle with $x=p_0$. Therefore, the only possible extra edges are between $X_3$ and $p_2$ and between $Y_3$ and $p_{\ell-2}$ (and this is only possible if $\ell\ge 3$). Let $X_4\subseteq X_3$ be those vertices of $X_3$ that are not adjacent to $p_2$ and let $Y_4\subseteq Y_3$ be those vertices of $Y_3$ that are not adjacent to $p_{\ell-2}$. If $|X_4|\ge Q$, then let $X'$ be a $Q$ element subset of $X_4$ and let $x'=p_0$; if $|X_4|< Q$, then let $X'$ be a $Q$ element subset of $X_3\setminus X_4$ and let $x'=p_2$. Similarly, if $|Y_4|\ge Q$, then let $Y'$ be a $Q$ element subset of $Y_4$ and let $y'=p_\ell$; if $|Y_4|< Q$, then let $Y'$ be a $Q$ element subset of $Y_3\setminus Y_4$ and let $y'=p_{\ell-2}$. Let $P'$ be the subpath of $P$ between $x'$ and $y'$. Note that this path has odd length (if $\ell=3$, it is possible that $x'=p_2$ and $y'=p_1$, but the length is still an odd number, namely one). Then $X'$, $Y'$, and $P'$ form an induced $\Gdoublebroom{|P'|}{Q}$ subgraph, a contradiction. \end{proof} As in Lemma~\ref{lemma:ramsey:nonbipartite}, we turn Lemma~\ref{lemma:ramsey:balancedbipartite0} into a statement on graph classes. \begin{lemma} \label{lemma:ramsey:balancedbipartite} Let~$\F$ be a hereditary family of bipartite graphs. Then at least one of the following is true. \begin{enumerate} \item There is an integer $b\ge 1$ such that every connected graph in $\F$ is $b$-thin. \item $\F$ is a superset of \Fpath. \item $\F$ is a superset of \Fclique. \item $\F$ is a superset of \Fbiclique. \item $\F$ is a superset of \Fsubdivstar. \item $\F$ is a superset of \Fdoublebroom{s} for some odd integer $s\ge 1$. \end{enumerate} \end{lemma} \begin{proof} Assuming that $\F$ is not a superset of $\Fpath$, there is an integer $\ell\ge 1$ such that $\Gpath{\ell}\not\in \F$. Then, as in the proof of Lemma~\ref{lemma:ramsey:nonbipartite}, we can assume that there is a $Q\ge 1$ such that $\F$ does not contain any of the following graphs: \begin{itemize} \item $\Gpath{Q}$ \item $\Gclique{Q}$, \item $\Gbiclique{Q}$, \item $\Gdoublebroom{s}{Q}$ for any odd integer $1\le s < \ell$, \item $\Gsubdivstar{Q}$. \end{itemize} For every $s\ge \ell$, $\Gdoublebroom{s}{Q}$ contains $\Gpath{\ell}$ as induced subgraph, hence none of the graphs listed in Lemma~\ref{lemma:ramsey:balancedbipartite0} is contained in $\F$. It follows that every connected bipartite graph in $\F$ is $b(Q)$-thin. \end{proof} Combining Lemmas~\ref{lemma:ramsey:nonbipartite} and \ref{lemma:ramsey:balancedbipartite}, the proof of Theorem~\ref{theorem:ramsey:1} follows. \restateramseysmallthin* \begin{proof} Let us apply first Lemma~\ref{lemma:ramsey:nonbipartite} on $\F$. If $\F$ is a superset of any of the families listed in items 2--6 of Lemma~\ref{lemma:ramsey:nonbipartite}, then we are done. Assume therefore that there is an integer $a\ge 1$ such that every connected nonbipartite graph in $\F$ has size at most $a$. Let $\F'$ contain every bipartite graph in $\F$ and let us apply Lemma~\ref{lemma:ramsey:balancedbipartite} on $\F'$. Again, if $\F$ is a superset of any of the families listed in items 2--6 of Lemma~\ref{lemma:ramsey:balancedbipartite}, then we are done. Therefore, we may assume that there is an integer $b\ge 1$ such that every connected graph in $\F'$ is $b$-thin. Observe now that if $C$ is a bipartite component of some $H\in \F$, then $C$ itself is in $\F$, as $\F$ is hereditary. This implies that $C$ is $b$-thin. Therefore, we have shown that $\F$ is $a$-small/$b$-thin: every nonbipartite component has at most $a$ vertices and every bipartite component is $b$-thin. \end{proof} \subsection{Proof of the dichotomy for packing problems} \label{subsection:packing:theorem} Using the algorithm of Section~\ref{subsection:subgraph:turing:upperbounds}, the hardness results for the basic families proved in Section~\ref{subsection:subgraph:turing:lowerbounds}, and the characterization proved in Section~\ref{sec:subgraph-ramsey}, we can prove Theorem~\ref{theorem:intro:packing}. \restateintropacking* \begin{proof} Let $\F$ be a hereditary class of graphs. If $\F$ is small/thin, then Theorem~\ref{theorem:kernel:packing} shows that \FSubgraphTest admits a polynomial many-one kernel. If $\F$ is not small/thin, then it is a superset of one of the families listed in Theorem~\ref{theorem:ramsey:1}. If $\F$ contains $\Ffountain{s}$ for some odd integer $s\ge 3$, then the \textup{WK[1]}-hard problem $\Ffountain{s}$-\Packing (Lemma~\ref{lemma:fountainpacking:wkhard}) can be reduced to \FPacking, hence \FPacking is also \textup{WK[1]}-hard. The situation is similar if $\F$ is a superset of any of the families listed in items 5--8 of Theorem~\ref{theorem:ramsey:1}. As explained in Section~\ref{section:outline:hardness}, if~$\F$ is a superset of \Fclique or \Fbiclique then \FPacking is \textup{W[1]}-hard and does not admit a polynomial many-one kernel unless \containment. Therefore, we have shown that if $\F$ is not small/thin, then \FPacking is \textup{WK[1]}-hard, \textup{W[1]}-hard, or \kPath-hard in all cases, completing the proof Theorem~\ref{theorem:intro:packing}. \end{proof} \section{Turing kernelization complexity of subgraph testing} In this section, we prove Theorem~\ref{theorem:intro:turingsubgraph} characterizing the hereditary classes $\F$ for which \FSubgraphTest admits a polynomial Turing kernel. In Section~\ref{subsection:subgraph:turing:upperbounds}, we invoke the marking algorithm developed in Section~\ref{sec:comp-repr-sets} to give a polynomial Turing kernel for splittable classes. In Section~\ref{subsection:subgraph:turing:lowerbounds}, we establish the basic hardness results that $\Fdiamondfan$-\SubgraphTest and $\Fsubdivtree{s}$-\SubgraphTest for any $s\ge 1$ are \textup{WK[1]}-hard. In Section~\ref{sec:subgraph-ramsey}, we prove Theorem~\ref{theorem:ramsey:separator} characterizing splittable hereditary classes. In Section~\ref{sec:proof-dich-subgr}, we put together all these results to complete the proof of Theorem~\ref{theorem:intro:turingsubgraph}. \subsection{Upper bounds} \label{subsection:subgraph:turing:upperbounds} To prove the positive part of Theorem~\ref{theorem:intro:turingsubgraph} for splittable classes, we try every possible image for the set $D$ that realizes the $(a,b,c,d)$-split and then invoke Lemma~\ref{lemma:kernel:generic} to obtain a bounded-size instance for each possible way of fixing the image. \restatesubgraphturing* \begin{proof} Let~$\F$ be a hereditary family that is $(a,b,c,d)$-splittable and let~$(G,H)$ be an instance of the \kFSubgraphTest problem. Recall that the parameter is~$k := |V(H)|$. We will construct a list~$\L$ of~$|V(G)|^{\Oh(1)}$ instances~$(G[X_1], H), \ldots, (G[X_t], H)$ of the \kFSubgraphTest problem, each of size polynomial in~$k$, such that~$H \subseteq G$ if and only if~$H \subseteq G[X_i]$ for some~$i \in [t]$. This easily implies the existence of a polynomial-size Turing kernel following Definition~\ref{definition:turing:kernelization}: we can query the oracle for the answer to each subinstance~$(G[X_i], H)$ of size polynomial in~$k$ and output the logical OR of the answers, which is the correct answer to the instance~$(G, H)$. The description in terms of a list of small instances highlights the fact that the Turing kernelization is non-adaptive and therefore amenable to parallelization. The kernelization algorithm starts by searching for a set~$D$ that realizes an~$(a,b,c,d)$-split of~$H$. As such a set~$D$ has size at most~$c$, which is a constant, can we try all possible sets~$\binom{V(G)}{\leq c}$ to determine whether there is one that realizes the split. Note that given a candidate set, it is easy to determine in polynomial time whether it realizes an~$(a,b,c,d)$-split of~$H$ or not. If the split cannot be realized then~$H \not \in \F$ and the instance does not satisfy the input requirements; we output \no. In the remainder we can work with a set~$D \subseteq V(G)$ of size at most~$c$ that realizes the split. We then proceed as follows. For each partial $H$-subgraph model~$\phi_i$ in~$G$ with domain~$D$, we invoke the algorithm of Lemma~\ref{lemma:kernel:generic} to the source graph~$G$, the query graph~$H$ with the separator~$D$, and the partial $H$-subgraph model~$\phi_0$. As~$D$ realizes an~$(a,b,c,d)$-split of~$H$, each connected component~$C$ of~$H - D$ that has size more than~$a$ is a $b$-thin bipartite graph in which the number of vertices whose closed neighborhood is not universal to~$N_H(C) \cap D$ is bounded by~$d$. As the domain of~$\phi_0$ is exactly~$D$, the lemma outputs a set~$X_i$ of size~$\Oh(k^{\Oh(a + b^2 + d)})$ such that~$G$ contains a full $H$-subgraph model that extends~$\phi_i$ if and only if~$G[X_i]$ contains a full $H$-subgraph model that extends~$\phi_i$. We add~$(G[X_i], H)$ to the list~$\L$. Since~$|D| \leq c$, which is a constant, the number of distinct partial $H$-subgraph models in~$G$ with domain~$D$ is~$\Oh(|V(G)|^c)$, which is polynomial in~$|V(G)|$ as~$c$ is a constant. As the computation of Lemma~\ref{lemma:kernel:generic} takes polynomial time for constant~$(a,b,c,d)$, the entire algorithm runs in polynomial time and the size of~$\L$ is bounded by a polynomial. To complete the proof it therefore suffices to show that if and only if~$H \subseteq G$ then~$H \subseteq G[X_i]$ for some~$i$. The reverse direction is trivial as~$G[X_i]$ is an induced subgraph of~$G$. For the forward direction, assume that~$\phi^*$ is a full~$H$-subgraph model in~$G$ and consider its restriction~$\phi^*|_D$ to the vertices of~$D$. Then~$\phi^*|_D$ is a partial $H$-subgraph model with domain~$D$, hence it occurred as a model~$\phi_i$ in our enumeration. As~$\phi^*$ is a full $H$-subgraph model in~$G$ that extends~$\phi_i = \phi^*|_D$, the guarantee of Lemma~\ref{lemma:kernel:generic} for set~$X_i$ ensures that~$G[X_i]$ contains a full $H$-subgraph model. Hence there is an index~$i$ such that~$H \subseteq G[X_i]$. By the argumentation given earlier, this concludes the proof. \end{proof} \subsection{Lower bounds} \label{subsection:subgraph:turing:lowerbounds} In this section we present the polynomial-parameter transformations that establish Theorem~\ref{theorem:subgraphtest:lowerbounds}, which we repeat here for the reader's convenience. \restatesubgraphlower* The proof of Theorem~\ref{theorem:subgraphtest:lowerbounds} follows from the following to lemmas, which show the \textup{WK[1]}-hardness of \Fdiamondfan-\SubgraphTest and \Fsubdivtree{s}-\SubgraphTest by polynomial-para\-meter transformations from \nRegularExactSetCover. \begin{lemma}\label{lemma:subgraph:diamondfan:wkhard} \Fdiamondfan-\SubgraphTest is \textup{WK[1]}-hard. \end{lemma} \begin{proof} We transform an instance~$(r, \S, U, n)$ of \nRegularExactSetCover in polynomial time into an equivalent instance~$(G, H)$ of \Fdiamondfan-\Packing with $H:= \Gdiamondfan{Q}$ and $Q:=n+1$. Note that the parameter is $k := |V(H)| =Q^2+Q+1\in \Oh(n^2)$ (see Figure~\ref{fig:basic}). Let $t:=n/r$, the number of sets in a solution. The host graph~$G$ is defined as follows. \begin{itemize} \item Graph~$G$ contains the vertex set~$U$ as an independent set. \item We introduce $Q-t$ copies of $K_{2,Q}$ and identify into a single vertex $z$ one degree-$Q$ vertex from each a copy. \item For every size-$r$ set~$S \in \S$, we introduce a vertex $v_S$ adjacent to all vertices representing members of~$S$. \item For every size-$r$ set $S\in S$, we introduce a set $X_S$ of $Q-r\ge Q-n\ge 0$ vertices that are adjacent to both $v_S$ and $z$. \end{itemize} Note that the degree of every $v_S$ is exactly $Q$. As the construction can be performed in polynomial time and produces an instance of the target problem whose parameter is suitably bounded, it remains to prove that the input and output instances are equivalent. \begin{claim} If~$(\S,U)$ has an exact cover, then~$H \subseteq G$. \end{claim} \begin{claimproof} Assume that~$(\S,U)$ has an exact cover~$S_1, \ldots, S_t \in \S$. For each~$S_i$ with~$i \in [t]$, the vertex $z$, vertex $v_S$, and the $Q$ neighbors of $v_S$ form a copy of $K_{2,Q}$. As the sets $S_i$'s are disjoint, these copies of $K_{2,Q}$ intersect only in $z$. Therefore, these copies together with the $Q-t$ copies introduced in the construction of $G$ form a $\Gdiamondfan{Q}$ subgraph centered at $z$ in $G$. \end{claimproof} Let $c$ be the unique vertex of $H$ having degree $Q^2$. Graph $H$ contains $Q$ degree-$Q$ vertices at distance two from $c$, let $W$ be the set of all these vertices. \begin{claim} \label{claim:diamondfan:hardness:goodmodels} If $\phi$ is a full subgraph model of $H$ in $G$, then $\phi(c)=z$. \end{claim} \begin{claimproof} Every vertex $v_S$ has degree $Q$ and every vertex in every $X_S$ has degree 2. The vertices of the copies of $K_{2,Q}$ introduced in the construction of $G$ (other than $z$) have degree two or $Q$. Therefore, $\phi(c)\neq z$ is only possible if $\phi(c)\in U$. The vertices of $W$ have to be mapped to vertices of degree at least $Q$ at distance exactly two from $\phi(c)$ in $G$. As $\phi(c)\in U$, the vertices of $G$ at distance two from $\phi(c)$ are the vertices of $U$, the degree-2 vertices in some $X_S$, and the degree-2 vertices of some $K_{2,Q}$ adjacent to $z$. Therefore, only at most $|U|=n<Q$ of them can have degree at least $Q$, a contradiction. \end{claimproof} \begin{claim} If~$H \subseteq G$, then~$(\S,U)$ has an exact cover. \end{claim} \begin{claimproof} Assume that $\phi$ is a full subgraph model of $H$ in $G$. By Claim~\ref{claim:diamondfan:hardness:goodmodels}, we have $\phi(c)=z$. The vertices of $W$ have to be mapped to vertices of $G$ at distance two from $z$. The candidates for these vertices are the vertices $v_S$ and the $Q-t$ degree-$Q$ vertices introduced in the $Q-t$ copies of $K_{2,Q}$. Therefore, at least $t$ vertices of $W$ are mapped to the $v_S$'s. Note that each vertex $v_S$ has degree exactly $Q$, hence the neighborhood of a vertex $w\in W$ is mapped bijectively to the neighborhood of $\phi(w)$. The neighborhoods of the vertices in $W$ are disjoint, hence the neighborhoods of the images should be disjoint as well. Therefore, the sets corresponding to the images of $W$ are disjoint, implying that there exists at least $t$ disjoint sets in $\S$. As these sets contain~at least $t \cdot r = n$ vertices in total, this is only possible if there are exactly $t$ of these sets and they cover all of~$U$, forming an exact cover. \end{claimproof} This concludes the proof of Lemma~\ref{lemma:subgraph:diamondfan:wkhard}. \end{proof} \begin{lemma}\label{lemma:subgraph:subdivtree:wkhard} For every $s\ge 1$, \Fsubdivtree{s}-\SubgraphTest is \textup{WK[1]}-hard. \end{lemma} \begin{proof} We transform an instance~$(r, \S, U, n)$ of \nRegularExactSetCover in polynomial time into an equivalent instance~$(G, H)$ of \Fsubdivtree{s}-\Packing with~$H:=\Gsubdivtree{s}{Q}$ and $Q:=n+2$. Note that the parameter is $k := |V(H)| =Q^2+Qs+1\in \Oh(n^2)$ (see Figure~\ref{fig:basic}). Let $t:=n/r$, the number of sets in a solution. The host graph~$G$ is defined as follows. \begin{itemize} \item Graph~$G$ contains the vertex set~$U$ as an independent set. \item We introduce a distinguished vertex $z$. \item We introduce $Q-t\ge Q-n\ge 0$ vertices $x_1$, $\dots$, $x_{Q-t}$, connect each $x_i$ to $z$ with a path of length $s$, and attach $Q$ pendant vertices to each $x_i$. \item For every size-$r$ set~$S \in \S$, we introduce a vertex $v_S$ adjacent to all vertices representing members of~$S$, connect $v_S$ and $z$ with a path of length $s$, and attach $Q-r\ge Q-n\ge 0$ pendant vertices to $v_S$. \end{itemize} As the construction can be performed in polynomial time and produces an instance of the target problem whose parameter is suitably bounded, it remains to prove that the input and output instances are equivalent. \begin{claim} If~$(\S,U)$ has an exact cover, then~$H \subseteq G$. \end{claim} \begin{claimproof} Assume that~$(\S,U)$ has an exact cover~$S_1, \ldots, S_t \in \S$. Consider the set $Z$ of vertices containing $v_{S_i}$ for $i \in [t]$ and $x_j$ for $j\in [Q-t]$. Each of these $Q$ vertices are connected to $z$ with a path of length $s$, and these paths intersect only in $z$. Each vertex of $Z$ has degree $Q+1$, that is, has $Q$ neighbors in addition to their neighbor on the path. Distinct vertices in $Z$ have disjoint neighborhoods: these neighbors are either degree-one (thus disjointness is trivial) or appear in $U$, where disjointness follows from the fact that $S_1$, $\dots$, $S_t$ are disjoint. Therefore, we have found a copy of $\Gsubdivtree{s}{Q}$ centered at $z$. \end{claimproof} Let $c$ be the unique vertex of $H$ having degree $Q$. \begin{claim} \label{claim:subdivtree:hardness:goodmodels} If $\phi$ is a full subgraph model of $H$ in $G$, then $\phi(c)=z$. \end{claim} \begin{claimproof} Clearly, $\phi(c)$ has degree at least $Q$. Vertices $x_i$ have degree $Q+1$, but they have only one neighbor with degree more than one, while every neighbor of $c$ has degree at least two. Vertices $v_S$ have degree $Q+1$, but they have at most $r+1<Q$ neighbors with degree more than one. Therefore, $\phi(c)=z$ is the only possibility.\end{claimproof} \begin{claim} If~$H \subseteq G$, then~$(\S,U)$ has an exact cover. \end{claim} \begin{claimproof} Assume that $\phi$ is a full subgraph model of $H$ in $G$. By Claim~\ref{claim:subdivtree:hardness:goodmodels}, we have $\phi(c)=z$. Graph $H$ contains $Q$ degree-$(Q+1)$ vertices at distance $s$ from $c$, let $W$ be the set of all these vertices. The vertices of $W$ have to be mapped to vertices of $G$ at distance exactly $s$ from $z$. The candidates for these vertices are the vertices $v_S$ and the $Q-t$ vertices $x_i$. Therefore, at least $t$ vertices of $W$ are mapped to the $v_S$'s. Note that each vertex $v_S$ has degree exactly $Q+1$, hence the neighborhood of a vertex $w\in W$ is mapped bijectively to the neighborhood of $\phi(w)$. The neighborhoods of the vertices in $W$ are disjoint, hence the neighborhood of the images should be disjoint as well. Therefore, the sets corresponding to the images of $W$ are disjoint, implying that there exists at least $t$ disjoint sets in $\S$. As these sets contain at least $t \cdot r = n$ vertices in total, this is only possible if there are exactly $t$ of these sets and they cover all of~$U$, forming an exact cover. \end{claimproof} This concludes the proof of Lemma~\ref{lemma:subgraph:diamondfan:wkhard}. \end{proof} \subsection{Combinatorial characterizations} \label{sec:subgraph-ramsey} To prove Theorem~\ref{theorem:ramsey:separator} characterizing hereditary classes that are not splittable, we need first the following auxiliary result. \begin{lemma} \label{lemma:ramsey:separator0} Let~$\F$ be a hereditary graph family. Then at least one of the following holds: \begin{enumerate} \item There is an integer $M\ge 1$ such that every $H\in \F$ has a set $S_0\subseteq V(H)$ of at most $M$ vertices such that every component of $H- S_0$ has vertex cover number at most $M$. \item $\F$ is a superset of at least one of \begin{itemize} \item $\Fpath$, \item $\Fclique$, \item $\Fbiclique$, \item $Q\cdot \Fsubdivstar$, or \item $Q\cdot \Ffountain{3}$. \end{itemize} \end{enumerate} \end{lemma} \begin{proof} Assuming that $\F$ does not contain any of the forbidden families, there is a $Q\ge 1$ such that $\F$ does not contain any of \begin{itemize} \item $\Gclique{Q}$, \item $\Gbiclique{Q}$, \item $Q\cdot \Gsubdivstar{Q}$, and \item $Q\cdot \Gfountain{3}{Q}$. \end{itemize} We set the following constants: \begin{align*} h&=P(Q,Q), \ \ \textup{(for the function $P$ in Corollary~\ref{theorem:path:ramsey2})}\\ M_1&=2Q(h+1)^2,\\ M_2&=\sum_{i=0}^{h}Q^i,\\ M&=\max\{M_1,M_2\}. \end{align*} Pick an arbitrary $H\in \F$ and compute a DFS tree for each component of $H$. The height of the DFS forest obtain this way is at most $h=P(Q,Q)$ by Corollary~\ref{theorem:path:ramsey2}, otherwise $H$ would contain $\Gpath{Q}$, $\Gclique{Q}$, or $\Gbiclique{Q}$ as induced subgraph. Let $L$ be the set of leaves of this forest and let $Z=V(H)\setminus L$. Let $Z^*$ be the those vertices of $Z$ that have at least $Q+1$ children in the DFS forest that belong to~$Z$. We claim that if there are $v_1$, $\dots$, $v_{2Q}$ vertices of $Z^*$ on the same level of the DFS forest, then $H$ contains $Q\cdot \Gfountain{s}{Q}$ or $Q\cdot \Gsubdivstar{Q}$ as an induced subgraph. As $\F$ is hereditary, it would follow that these graphs are in $H$, a contradiction. Let $x_{i,1}$, $\dots$, $x_{i,Q+1}$ be children of $v_i$ in $Z$. As they are not leaves, $x_{i,j}$ has a child $y_{i,j}$. By the properties of the DFS forest, there is no edge between $\{x_{i,j},y_{i,j}\}$ and $\{x_{i,j'},y_{i,j'}\}$ for any $j\neq j'$. If there is an edge between $v_i$ and $y_{i,j'}$, then the triangle $\{v_i,x_{i,j'},y_{i,j'}\}$ and the vertices $x_{i,j}$, $1 \le j \le Q+1$, $j\neq j'$ form an induced $\Gfountain{3}{Q}$. If there is no edge between $v_i$ and $y_{i,j}$ for any $1\le j \le Q$, then $\{v_i,x_{i,1},\dots,x_{i,Q},y_{i,1},\dots, y_{i,Q}\}$ induces a $\Gsubdivstar{Q}$. Thus for every $1\le i \le 2Q$, we get an induced $\Gfountain{3}{Q}$ or $\Gsubdivstar{Q}$ on $v_i$, its children, and its grandchildren. As all the $v_i$'s are on the same level of the DFS forest, there is no edge between these $2Q$ graphs. Thus we get either $Q$ independent copies of $\Gfountain{3}{Q}$ or $Q$ independent copies of $\Gsubdivstar{Q}$, a contradiction. We have proved that each level of the DFS forest contains less than $2Q$ vertices of $Z^*$, implying that $|Z^*|\le 2Q(h+1)$. Let $S_0$ contain every vertex of $Z^*$ and every ancestor of every vertex in $Z^*$: as the DFS forest has height at most $h$ (that is, at most $h+1$ levels), we have $|S_0|\le (h+1)|Z^*|=2Q(h+1)^2=M_1$. Observe that if $u$ and $v$ are adjacent vertices in $H- S_0$ and $u$ is an ancestor of $v$ in the DFS forest, then the unique $u-v$ path of the DFS tree is disjoint from $S_0$: if any vertex of this path is in $S_0$, then $u$ itself is in $S_0$. Therefore, if $C$ is the set of vertices of a component of $H- S_0$, then $C$ induces a connected subtree of the DFS forest. As every vertex of $C\cap Z$ has at most $Q$ children in $Z$ (otherwise it would be in $S_0$) and the tree has height at most $h$, we have that $|C\cap Z|\le \sum_{i=0}^{h}Q^i=M_2$. Observe furthermore that $C\cap Z$ is a vertex cover of $H[C]$: if there is an edge between $u,v\in C\setminus Z$, then it is an edge between two leaves of the DFS forest. Therefore, $S_0$ is a set of at most $M_1\le M$ vertices such that every component of $H- S_0$ has vertex cover number at most $M_2\le M$, what we had to show. \end{proof} Now we are ready to prove Theorem~\ref{theorem:ramsey:separator}. \restateramseysplittable* \begin{proof} By Lemma~\ref{lemma:ramsey:separator0}, the assumption that $\F$ is not the superset of the first 5 families listed in the lemma implies that there is an $M\ge 1$ such that every $H\in \F$ has a set $S_0\subseteq V(H)$ of at most~$M$ vertices such that every component of $H- S_0$ has vertex cover number at most $M$. Assuming that $\F$ does not contain any of the forbidden families, there is a $Q\ge 1$ such that $\F$ does not contain \begin{itemize} \item $\Gpath{Q}$, \item $\Gclique{Q}$, \item $\Gbiclique{Q}$, \item $Q\cdot \Gsubdivstar{Q}$, \item $Q\cdot \Gfountain{s}{Q}$ for any integer $3\le s \le Q+2$, \item $Q\cdot \Goperahouse{s}$ for any odd integer $1 \le s \le Q$, \item $Q\cdot \Gdoublebroom{s}$ for any odd integer $1 \le s \le Q$, \item $Q\cdot \Glongfountain{s}{t}$ for any odd integer $3\le s \le Q+2$ and integer $1\le t \le Q$, \item $\Gsubdivtree{s}{Q}$ for any $1\le s \le Q$, \item $\Gdiamondfan{Q}$. \end{itemize} (The argument why there is such a finite $Q$ is the same as in the proof of Lemma~\ref{lemma:ramsey:nonbipartite}.) We set the following constants: \begin{align*} a&=a(Q),\ \ \textup{(for the function $a(Q)$ in Lemma~\ref{lemma:ramsey:nonbipartite0})}\\ b&=b(Q),\ \ \textup{(for the function $b(Q)$ in Lemma~\ref{lemma:ramsey:balancedbipartite0})}\\ k&=Q(Q^2+2Q)+Q(Q+1)+QM+Q^2M,\\ c&=(k+1)M,\\ d&=Q\cdot 2^{b+M}+b. \end{align*} Select an arbitrary $H\in F$ and let $S_0$ be given by Lemma~\ref{lemma:ramsey:separator0}. We say that component is $C$ of $H- S_0$ is {\em good} if it has at most $a$ vertices or it is a $b$-thin bipartite graph with at most $d$ vertices whose neighborhood in~$H[C]$ is not universal to $N(C)\cap S_0$. Otherwise, we say that the component is {\em bad,} which means that at least one of the following holds: \begin{enumerate} \item $C$ is nonbipartite and has more than $a$ vertices. \item $C$ is bipartite, but not $b$-thin. \item $C$ is bipartite, is $b$-thin, but has more than $d$ vertices whose closed neighborhood in~$H[C]$ is not universal to $N(C)\cap S_0$. \end{enumerate} The following claim bounds the number of bad components. Then we show how to extend $S_0$ to destroy the bad components. \begin{claim}\label{claim:boundbad} $H - S_0$ has at most $k$ bad components. \end{claim} \begin{proof} Let us bound first the number nonbipartite components that have at least $a$ vertices. As $H$ does not contain $\Gpath{Q}$ or $\Gclique{Q}$ as an induced subgraph, Lemma~\ref{lemma:ramsey:nonbipartite0} implies that each such nonbipartite component contains either $\Gfountain{s}{Q}$ for some odd integer $3\le s\le Q+2$, $\Glongfountain{s}{t}{Q}$ some odd integer $3\le s\le Q+2$ and integer $1\le t\le Q$, or $\Goperahouse{s}{Q}$ for some odd integer $1\le s\le Q$ as induced subgraph. That is, each nonbipartite component of size at least $a$ contains one of these $Q^2+2Q$ graphs, hence if there are $Q(Q^2+2Q)$ such components, one of these graphs appears at least $Q$ times, contradicting our assumptions. Next we bound the number of bipartite components that are not $b$-thin. As $\Gpath{Q}$, $\Gclique{Q}$, and $\Gbiclique{Q}$ do not appear in $H$ as induced subgraph, Lemma~\ref{lemma:ramsey:balancedbipartite0} implies that each such bipartite component contains $\Gsubdivstar{Q}$ or $\Gdoublebroom{s}{Q}$ for some $1\le s \le Q$ as induced subgraph. Therefore, if there are $Q(Q+1)$ bipartite non-thin components, one of these graphs appears at least $Q$ times, contradicting our assumptions. Finally, we bound the number of components that are bipartite, $b$-thin, but have more than $d$ vertices whose closed neighborhood is not universal to $N(C)\cap S_0$. In particular, there is a set $X_0$ of at least $d-b$ such vertices in the larger bipartite class. Each vertex in the larger class of $C$ has degree at most $b+|S_0|\le b+M$ in~$H$, hence we can partition $X_0$ into at most $2^{b+M}$ classes according to their neighborhood. Therefore, there is a set $X\subseteq X_0$ of at least $|X_0|/2^{b+M}\ge (d-b)/2^{b+M}=Q$ such vertices in the larger side of $C$ whose neighborhoods are the same. There are two possibilities why the closed neighborhoods of the vertices in $X$ are not universal to $N(C)\cap S_0$: either some neighbor of $X$ is not universal to $N(C)\cap S_0$, or the vertices of $X$ are not universal to $N(C)\cap S_0$. \begin{figure}[t] \begin{center} {\small \svg{\linewidth}{diamondfan2}} \caption{Proof of Theorem~\ref{theorem:ramsey:separator}, Claim~\ref{claim:boundbad}. Finding (a) a \Gdiamondfan{4} or (b) a \Gsubdivtree{5}{4} in four bad components.}\label{fig:dfan} \end{center} \end{figure} The first case is when every vertex of $X$ is universal to $N(C)\cap S_0$, but they have a neighbor $x\in C$ that is not adjacent to some $z\in N(C)\cap S_0$. If there are at least $Q|S_0|$ components where this case happens, then there are at least $Q$ of them for which this case happens with the same $z\in S_0$. Then the corresponding vertices $x$ and sets $X$ form a $\Gdiamondfan{Q}$ (see Figure~\ref{fig:dfan}(a)). The second case is that the vertices in $X$ are not adjacent to some vertex $z\in N(C)\cap S_0$. The fact that $z$ is in $N(C)$ implies that $z$ has a neighbor $y\in C$. Let us find a shortest path in $H[C]$ from each vertex of $X$ to $y$. As every vertex in $X$ has the same set of neighbors, we may assume that the second vertex is the same for each of these $|X|$ paths. In other words, there is a $v\in C$ and a $v-y$ path $P$ such that $v$ is adjacent to every vertex in $X$ and, for every $x\in X$, the path $xP$ is a shortest $x-y$ path. This implies that $P$ is an induced path and $x$ is not adjacent to any vertex of $P$ except $v$. As $H$ does not contain $\Gpath{Q}$ as induced subgraph, we also get that $|P|<Q$. If there are at least $Q^2|S_0|$ components where this subcase happens, then there are $Q$ of them for which this subcase happens with the same $z\in S_0$ and the length of $P$ is the same integer $1\le s \le Q$. Then the corresponding paths $P$ and sets $X$ form a $\Gsubdivtree{s}{Q}$ (see Figure~\ref{fig:dfan}(b)). Summing up all cases, we get that $H- S_0$ has at most $k=Q(Q^2+2Q)+Q(Q+1)+QM+Q^2M$ bad components. \cqed\end{proof} We have shown that at most $k$ of the components of $H- S_0$ are bad components. Then we destroy these bad components by extending $S_0$ by a minimum vertex cover of each bad component $C$; let $S$ be the resulting set of vertices. We have seen that (Lemma~\ref{lemma:ramsey:separator0}) that each component $C$ has a vertex cover of size $M$, hence $|S|\le |S_0|+kM\le (k+1)M=c$. Observe that if $C$ is a bad component of $H- S_0$, then every vertex of $C\setminus S$ is an isolated vertex of $H- S$. Moreover, the good components of $H- S_0$ are unaffected by extending $S_0$ to $S$ and it is also true that $N(C)\cap S_0=N(C)\cap S$ for every good component $C$ of $H- S_0$. Therefore, every component $C$ of $H- S$ either has size at most $a$ or it is a $b$-thin bipartite graph having at most $d$ vertices whose closed neighborhood is not universal to $N(C)\cap S$. As this is true for every $H\in \F$, we have shown that $\F$ is splittable. \end{proof} \subsection{Proof of the dichotomy for subgraph testing} \label{sec:proof-dich-subgr} Using the algorithm of Section~\ref{subsection:subgraph:turing:upperbounds}, the hardness results for the basic families proved in Section~\ref{subsection:subgraph:turing:lowerbounds}, the characterization proved in Section~\ref{sec:subgraph-ramsey}, and the hardness results for \FPacking obtained in Section~\ref{subsection:packing:lowerbounds}, we can prove Theorem~\ref{theorem:intro:turingsubgraph}. \restatemainsubgraph* \begin{proof} Let $\F$ be a hereditary class of graphs. If $\F$ is splittable, then Theorem~\ref{theorem:kernel:subgraph} shows that \FSubgraphTest admits a polynomial Turing kernel. Otherwise, Lemma~\ref{theorem:fpacking:pvsnp} gives a list of classes such that one of these classes is fully contained in $\F$. If $\F$ is a superset of $\Fpath$, then \FSubgraphTest is clearly \textsc{Long Path}-hard. If $\F$ is superset of $\Fclique$ or~$\Fbiclique$, then \FSubgraphTest W[1]-hard~\cite{DowneyF13,Lin14}. If $\F$ contains $n\cdot \Gsubdivstar{n}$ (that is, the disjoint union of $n$ copies of $\Gsubdivstar{n}$, then, as $\F$ is hereditary, it also contains $t\cdot \Gsubdivstar{n}$ for every $t,n\ge 1$. This means that an instance $(G,H,t)$ of \Packing with $H=\Gsubdivstar{n}$ can be expressed as an instance $(G,H')$ of $\FSubgraphTest$ with $H'=t\cdot H$. Therfore, $\Fsubdivstar$-\Packing, which was shown to be \textup{WK[1]}-hard in Theorem~\ref{claim:subdivstar:hardness:goodmodels}, can be reduced to $\FSubgraphTest$, implying that the latter problem is \textup{WK[1]}-hard as well. The situation is similar for items 5-8 in Theorem~\ref{theorem:ramsey:separator}: a \textup{WK[1]}-hard packing problem can be reduced to \FSubgraphTest. If $\F$ contains $\Fsubdivtree{s}$ for some $s\ge 1$, then $\FSubgraphTest$ is \textup{WK[1]}-hard by Theorem~\ref{lemma:subgraph:subdivtree:wkhard}. Similarly, if $\F$ contains $\Fdiamondfan$, then $\FSubgraphTest$ is \textup{WK[1]}-hard by Theorem~\ref{lemma:subgraph:diamondfan:wkhard}. Therefore, we have shown that if $\F$ is not splittable, then it is \textsc{Long Path}- or \textup{WK[1]}-hard in all cases, completing the proof of Theorem~\ref{theorem:intro:turingsubgraph}. \end{proof} \section{Many-one kernelization complexity of subgraph testing} For the case of many-one kernelization, we cannot determine the kernelization complexity for all hereditary graph families \F. The reason is that the complexity landscape is very diverse in this case, which we show by presenting several surprising kernelization upper and lower bounds. Obviously, since many-one kernelization is more restrictive than Turing kernelization, the negative results from Turing kernelization for \kFSubgraphTest carry over. However, we will see that for some of the families~\F where polynomial-size Turing kernels exist there is no polynomial-size many-one kernel unless \containment. \subsection{Lower bounds} \label{subsection:subgraph:karp:lowerbounds} \subsubsection{Kernelization lower bounds by OR-cross-composition} The kernelization lower bounds presented until this point employed polynomial-parameter transformations, as they serve simultaneously as \textup{WK[1]}-hardness proofs (obtaining Turing kernelization lower bounds under the assumption that no \textup{WK[1]}-hard problem admits a polynomial Turing kernel) and transformations from incompressible problems (obtaining many-one kernelization lower bounds under the assumption that \ncontainment). The lower bounds in Section~\ref{subsection:subgraph:karp:lowerbounds} have to employ different machinery as they apply only to many-one kernelization (Theorem~\ref{theorem:kernel:subgraph} provides polynomial-size Turing kernels for these problems). We use the technique of OR-cross-composition~\cite{BodlaenderJK14}, which builds on earlier results by Bodlaender et al.~\cite{BodlaenderDFH09} and Fortnow and Santhanam~\cite{FortnowS11}. \begin{definition} \label{definition:poly:eqv:relation} An equivalence relation~\eqvr on $\Sigma^*$ is called a \emph{polynomial equivalence relation} if the following two conditions hold: \begin{enumerate} \item There is an algorithm that given two strings~$x,y \in \Sigma^*$ decides whether~$x$ and~$y$ belong to the same equivalence class in~$(|x| + |y|)^{\Oh(1)}$ time. \item For any finite set~$S \subseteq \Sigma^*$ the equivalence relation~$\eqvr$ partitions the elements of~$S$ into at most~$(\max _{x \in S} |x|)^{\Oh(1)}$ classes. \end{enumerate} \end{definition} \begin{definition} \label{definition:cross:composition} Let~$L \subseteq \Sigma^*$ be a set and let~$\Q \subseteq \Sigma^* \times \mathbb{N}$ be a parameterized problem. We say that~$L$ \emph{OR-cross-composes} into~$\Q$ if there is a polynomial equivalence relation~$\eqvr$ and an algorithm that, given~$r$ strings~$x_1, x_2, \ldots, x_r \in \Sigma^*$ belonging to the same equivalence class of~$\eqvr$, computes an instance~$(x^*,k^*) \in \Sigma^* \times \mathbb{N}$ in time polynomial in~$\sum _{i \in [r]} |x_i|$ such that: \begin{enumerate} \item~$(x^*, k^*) \in \Q \Leftrightarrow x_i \in L$ for some~$i \in [r]$, \item~$k^*$ is bounded by a polynomial in~$\max _{i \in [r]} |x_i|+\log r$. \end{enumerate} \end{definition} \begin{theorem}[\cite{BodlaenderJK14}] \label{theorem:cross:composition:no:kernel} If a set~$L \subseteq \Sigma^*$ is NP-hard under many-one reductions and~$L$ OR-cross-composes into the parameterized problem~$\Q$, then there is no polynomial many-one kernel for~$\Q$ unless \containment. \end{theorem} \subsubsection{Canonical template graphs for packing problems} Before presenting the two OR-cross-compositions that prove the two main lower bounds for subgraph testing problems, we discuss the common general idea behind the constructions and give some preliminary lemmas. We will give a superpolynomial many-one kernelization lower bound for detecting a subdivided star together many vertex-disjoint triangles as a subgraph (Theorem~\ref{theorem:karp:lowerbound:onesubstar:manytriangles}), and for testing two subdivided stars together with many vertex-disjoint $P_3$'s as a subgraph (Theorem~\ref{theorem:karp:lowerbound:twosubstar:manyps}; recall that~$P_3$ is the length-two path on three vertices). Hence in both cases the task is to detect a constant number (one or two) large, constant-radius subgraphs together with many constant-size subgraphs. For an OR-cross-composition, we have to embed the logical OR of many instances of an NP-hard problem into a single instance of the target problem with a small value of the parameter. To obtain this OR behavior we make use of the fact that both for~$K_3$-packing and for~$P_3$-packing there are \emph{canonical template graphs} containing NP-complete instances: there are polynomial-time constructable graph families~$\G^{K_3}$ and~$\G^{P_3}$ such that all size-$n$ instances of the NP-complete \XTC can be reduced to~$K_3$ packing (respectively~$P_3$ packing) instances on induced subgraphs of the $n$-th member of the family~$\F^{K_3}$ (resp.~$\F^{P_3}$). \begin{figure}[t] \begin{center} \subfigure[The triangle gadget.]{\label{fig:trianglegadget} \begin{tikzpicture}[thick,>=stealth,x=0.5cm,y=0.5cm] \foreach \i \j \k in {a/0/0,b/3/0,c/6/0,al/-1/1,ar/1/1,bl/2/1,br/4/1,cl/5/1,cr/7/1,at/0/3,bt/3/2,ct/6/3} \node[fill=black,circle,inner sep=0,minimum size=0.2cm] (\i) at (\j, \k) {}; \foreach \i \j in {a/al,a/ar,at/al,at/ar,b/bl,b/br,bt/bl,bt/br,c/cl,c/cr,ct/cl,ct/cr,at/bt,bt/ct,ct/at,al/ar,bl/br,cl/cr} \draw (\i) -- (\j); \node[fill=black,rectangle,minimum size=0.2cm, inner sep=0,label=below:$x$,draw] (adesc) at ($(a)$) {}; \node[fill=black,rectangle,minimum size=0.2cm, inner sep=0,label=below:$y$,draw] (bdesc) at ($(b)$) {}; \node[fill=black,rectangle,minimum size=0.2cm, inner sep=0,label=below:$z$,draw] (cdesc) at ($(c)$) {}; \end{tikzpicture} } \subfigure[The~$P_3$ gadget.]{\label{fig:pathgadget} \begin{tikzpicture}[thick,>=stealth,x=0.5cm,y=0.5cm] \foreach \i \j \k in {a/0/0,b/3/0,c/6/0,al/0/1,ar/1/1,bl/3/1,br/4/1,cl/6/1,cr/7/1,at/0/2,bt/3/2,ct/6/2} \node[fill=black,circle,inner sep=0,minimum size=0.2cm] (\i) at (\j, \k) {}; \foreach \i \j in {a/al,at/al,b/bl,bt/bl,c/cl,ct/cl,at/bt,bt/ct,al/ar,bl/br,cl/cr} \draw (\i) -- (\j); \node[fill=black,rectangle,minimum size=0.2cm, inner sep=0,label=below:$x$,draw] (adesc) at ($(a)$) {}; \node[fill=black,rectangle,minimum size=0.2cm, inner sep=0,label=below:$y$,draw] (bdesc) at ($(b)$) {}; \node[fill=black,rectangle,minimum size=0.2cm, inner sep=0,label=below:$z$,draw] (cdesc) at ($(c)$) {}; \end{tikzpicture} } \caption{Gadgets for the canonical template graph families~$\G^{K_3}$ and~$\G^{P_3}$. The vertices represented by circles are private to the gadget, while the three vertices~$x,y,z$ represented by squares are shared by other gadgets.} \end{center} \end{figure} \begin{definition} \label{definition:canonicalgraphs} For each positive integer~$n$, the $n$-th graph~$\G^{K_3}_n$ in the sequence of \emph{canonical template graphs for triangle packing}~$\G^{K_3}$ is obtained as follows: \begin{itemize} \item Start with an independent set~$U$ of size~$n$. \item For each set~$\{x,y,z\} \in \binom{U}{3}$, add a copy of the nine private vertices in the triangle gadget of Figure~\ref{fig:trianglegadget} to the graph and connect them to the vertices~$x,y,z \in U$ as in Figure~\ref{fig:trianglegadget}. \end{itemize} The $n$-th graph~$\G^{P_3}_n$ in the sequence of \emph{canonical template graphs for~$P_3$ packing} is obtained in the same way, using the path gadget of Figure~\ref{fig:pathgadget} instead of the triangle gadget. \end{definition} From this definition it easily follows that for each~$n$, the graphs~$\G^{K_3}_n$ and~$\G^{P_3}_n$ can be constructed in time polynomial in~$n$ and have size exactly~$n + 9\binom{n}{3}$. The graph families are called canonical because the following problems are NP-complete. \problemdef{\TrianglePackingCanonical} {An integer~$n$ encoded in unary and a subset~$S$ of the vertices of~$\G^{K_3}_n$.} {Can the vertices of~$\G^{K_3}_n[S]$ be partitioned into triangles?} \problemdef{\PathPackingCanonical} {An integer~$n$ encoded in unary and a subset~$S$ of the vertices of~$\G^{P_3}_n$.} {Can the vertices of~$\G^{P_3}_n[S]$ be partitioned into~$P_3$'s?} \begin{lemma} \label{lemma:canonicalpacking:npcomplete} \TrianglePackingCanonical and \PathPackingCanonical are NP-complete. \end{lemma} \begin{proof} Membership in NP is trivial. Completeness follows from the fact that the existing NP-completeness reductions for triangle packing and~$P_3$ packing, which follow from more general results by Kirkpatrick and Hell~\cite{KirkpatrickH78}, construct induced subgraphs of the canonical graphs. In particular, in these two specific cases their NP-completeness proof reduces an instance of \XTC to instances of triangle packing or~$P_3$ packing. An instance of \XTC consists of a universe~$U$ among with a collection~$T \subseteq \binom{U}{3}$ of size-three subsets of the universe, and asks whether there is a subset~$T' \subseteq T$ such that each element of~$U$ is contained in exactly one set of~$T'$. To reduce such an instance~$(U,T)$ of \XTC to triangle packing or~$P_3$ packing, it suffices to perform the construction of Definition~\ref{definition:canonicalgraphs} but to only make gadgets for the size-three subsets of~$U$ that appear in~$T'$. Consequently, we can obtain instances of \TrianglePackingCanonical and \PathPackingCanonical that are equivalent to~$(U,T$) by letting~$S$ contain the vertices~$U$ together with those of the gadgets that would be created in the reduction. \end{proof} \subsubsection{One subdivided star and many triangles} Before presenting the technical details, we describe the main intuition of the construction. We OR-cross-compose a sequence of instances of the NP-complete \TrianglePackingCanonical problem and choose a polynomial equivalence relation that enforces that all instances we are working with share the same value of~$n$ and have the same size set~$S$. The instance of the \kFSubgraphTest problem that we create contains the canonical graph~$\G^{K_3}_n$ for instances of size~$n$. For each vertex of~$\G^{K_3}_n$ we will add a private triangle of two new vertices that cannot be used in any other triangles; this private triangle will therefore be a \emph{cheap} way to get a triangle subgraph, as they consume only one vertex of~$\G^{K_3}_n$. The only other triangles in~$G$ will be within~$\G^{K_3}_n$, which are \emph{expensive} since they consume three vertices of~$\G^{K_3}_n$. We encode the input instances through the following mechanism. For each input instance~$i$, which is completely characterized by the set~$S_i$ for which it asks whether~$\G^{K_3}_n[S_i]$ can be partitioned into triangles, we add one possibility of realizing a large subdivided star to the host graph~$G$, such that realizing the subdivided star in this way blocks the cheap triangles using~$S_i$. As~$H$ will consist of one large subdivided star together with many disjoint triangles, to find an $H$-subgraph one must find a subdivided star together with many disjoint triangles. We will choose the number of triangles in such a way that, after a realization of a subdivided star is chosen that blocks the cheap triangles of~$S_i$, it will be optimal to take the cheap triangles for all vertices of~$\G^{K_3}_n - S_i$, effectively eliminating these vertices. The number of required triangles will be set such that, after taking these cheap triangles, one has to partition the remaining graph~$\G^{K_3}_n[S_i]$ into triangles to get a sufficient number. Armed with this intuition we present the formal proof. \restatekarponesubstarmanytriangles* \begin{proof} We will prove that \TrianglePackingCanonical OR-cross-composes into the problem of testing whether a graph~$G$ contains a subgraph of the form~$H \in \Fsubdivstar + \ell \cdot K_3$. Following the definition of OR-cross-composition we first define a polynomial equivalence relation~\eqvr on strings. We let all strings that do not encode a valid instance of the problem be equivalent, and we let valid instances be equivalent if they agree on the value of~$n$ and the size of the deleted set~$S$. As the size of~$\G^{K_3}_n$ is polynomial in~$n$, for each value of~$n$ there are only polynomially many choices for the size of the set~$S$ which easily implies that this is a polynomial equivalence relation. In the remainder, it suffices to show how to OR-cross-compose a sequence of instances that is equivalent under~\eqvr. If the strings do not encode valid instances, then we output a constant-size \no-instance as the result of the OR-cross-composition. From now on we may therefore assume that~$x_1, \ldots, x_r$ is a series of strings encoding valid instances~$(n, S_1), \ldots, (n, S_r)$ whose sets~$S_i$ all have the same size~$m$. If~$n < 10$ or~$m < 10$, then we can solve all instances in constant time and output the appropriate answer; hence we may assume~$n,m \geq 10$. Let~$\G := \G^{K_3}_n$ be the canonical graph for the current value of~$n$. We construct graphs~$G$ and~$H$ such that~$H \subseteq G$ if and only if there is a \yes-instance among the inputs. \begin{enumerate} \item We initialize~$G$ as a copy of~$\G$. For every vertex~$v \in V(\G)$, we add a \emph{dummy} vertex~$v'$ and an \emph{activator} vertex~$v''$ and turn~$\{v,v',v''\}$ into a triangle. \item For each instance number~$i \in [r]$, we create an \emph{instance selector}~$u_i$ and make it adjacent to~$\{v'' \mid v \in S_i\}$. \end{enumerate} This concludes the description of~$G$. \begin{observation}\label{observation:triangle:lowerbound} Each triangle in~$G$ contains at least one vertex of~$V(\G)$. Each triangle in~$G$ that contains less than three vertices of~$V(\G)$ contains one vertex~$v \in V(\G)$ and the corresponding dummy vertex~$v'$ and activator vertex~$v''$. \end{observation} Define~$t := |V(\G)| - m + m/3$. We let~$H'$ be a subdivided star with~$m$ leaves and pick~$H := H + t \cdot K_3$ to complete the description of the \kFSubgraphTest instance~$(G,H)$. Recall that the parameter to the problem is the value~$k := |V(H)| = (2m + 1) + 3 t = (2m+1) + 3(|V(\G)| - m + m/3)$. As~$\G = \G^{K_3}_n$ has size polynomial in~$n$ and~$m$ measures the size of a vertex subset of~$\G$, it follows that~$k$ is polynomial in~$n$ and therefore in the size of the largest input instance. Hence~$k$ is bounded appropriately for an OR-cross-composition. It is easy to see that the instance~$(G,H)$ can be constructed in polynomial time. It remains to prove that it acts as the logical OR of the input instances. To this end, we first establish the following claim. \begin{claim} \label{claim:triangle:tightbudget} Any packing of~$t$ vertex-disjoint triangles in~$G$ that uses at most~$|V(\G)| - m$ activator vertices, uses exactly~$|V(\G)| - m$ activator vertices and all vertices of~$V(\G)$. \end{claim} \begin{claimproof} First consider a packing using exactly~$|V(\G)| - m$ activator vertices. By Observation~\ref{observation:triangle:lowerbound}, every triangle uses at least one vertex of~$V(\G)$. For each activator vertex used, we can get a cheap triangle containing only one vertex of~$V(\G)$. The remaining~$t - (|V(\G)| - m) = m/3$ triangles each use three vertices from~$V(\G)$. Since~$m$ vertices of~$V(\G)$ are left after discarding those used in cheap triangles, these remaining~$m$ vertices must all be used in a triangle to get the additional~$m/3$ triangles needed to find~$t$ in total. Using similar arguments it is easy to show that if fewer than~$|V(\G)| - m$ activator vertices are used, one cannot obtain~$t$ in total. The claim follows. \end{claimproof} \begin{claim} $H \subseteq G$ if and only if there is an~$i \in [r]$ such that~$\G^{K_3}_n[S_i]$ can be partitioned into triangles. \end{claim} \begin{claimproof} ($\Leftarrow$) We first prove the easy reverse direction. Assume that~$\G^{K_3}_n[S_i]$ can be partitioned into triangles. As~$|S_i| = m$, this implies that~$G^{K_3}_n[S_i]$ contains~$m/3$ vertex-disjoint triangles. We extend this packing of triangles to an $H$-subgraph in~$G$ as follows. For each vertex~$v \in V(\G) \setminus S_i$, add the triangle~$\{v,v',v''\}$ to the subgraph model, increasing the number of triangles to~$(|V(\G)| - m) + m/3$ in total. Observe that none of these triangles contains vertices~$\{v'' \mid v \in S_i\}$. Hence we can realize a subdivided star that is disjoint from this packing of triangles by centering the star at vertex~$u_i$ and using the~$2m$ vertices~$\{v', v'' \mid v \in S_i\}$ for the~$m$ legs of the star. Using the construction of~$G$, it is easy to verify that all the required edges are present. As the model of the subdivided star is disjoint from the model of the triangles, we find~$H$ as a subgraph in~$G$. ($\Rightarrow$) For the forward direction, assume that~$\phi$ is a full $H$-subgraph model in~$G$. Let~$c \in V(H)$ be the center of the subdivided star, and let~$\phi'$ be the model of the subdivided star obtained by restricting~$\phi$. We will first establish that~$\phi'(c) = \phi(c) \in \{u_1, \ldots, u_r\}$. \begin{itemize} \item As dummy vertices of the form~$v'$ for~$v \in V(\G)$ have degree two in~$G$, they cannot model the center of a star of degree~$m \geq 10$. \item Suppose that~$\phi'(c) \in V(\G)$. Then the images of the~$m$ subdivider vertices of the star lie in~$N_G(\phi'(c))$. Observe that all neighbors of~$v \in V(\G)$ except~$v', v''$ belong to~$V(\G)$. Hence at least~$m - 2$ vertices of~$V(\G)$ are used in the model of the subdivided star. By Observation~\ref{observation:triangle:lowerbound}, each triangle in~$G$ uses at least one vertex of~$V(\G)$. As the models of the triangles are disjoint from the model of the subdivided star and there are at most~$|V(\G)| - (m - 2)$ vertices of~$V(\G)$ not used by the star, there can be at most~$|V(\G)| - (m - 2) < t$ triangles in the subgraph model. Hence~$\phi$ does not model~$H$; a contradiction. \item Suppose that~$\phi'(c)$ is an activator vertex of the form~$v''$ for~$v \in V(\G)$. Observe that the activator vertex~$v''$ has the unique $G$-neighbor~$v$ in the set~$V(\G)$, it has the dummy vertex~$v'$ as a neighbor, and it can have many instance selector vertices as $G$-neighbors. The vertices~$v$ and~$v'$ in~$G$ cannot be the images of vertices of two \emph{different} legs of the subdivided star: if~$v'$ is used in model~$\phi'$, then it either models a subdivider vertex (which means that its unique neighbor~$v$ distinct from~$\phi'(c)$ models the degree-one leaf of that leg of the star), or~$v'$ models a degree-one leaf which means that its only other neighbor~$v$ must model the subdivider vertex of the same leg of the star. Finally, if~$v'$ is not used in the model, then~$\{v, v'\}$ models part of (at most) one leg of the star. From these observations, we deduce the following. Since the only $G$-neighbors of the activator vertex~$\phi'(c) = v''$ are~$v$,~$v'$, and instance selector vertices, at least~$m - 1$ subdivider vertices of the star are mapped to instance selectors by~$\phi'$. As the only neighbors of instance selector vertices in~$G$ are activator vertices, for each instance selector that is used by~$\phi'$, an activator vertex distinct from~$\phi'(c) = v''$ must also be used by~$\phi'$. Hence we find: if~$v \in V(\G)$ is not used in model~$\phi'$, then~$m$ instance selector vertices are used, implying that~$m$ activator vertices distinct from~$\phi'(c)$ are also used, giving a total number of~$m + 1$ activator vertices if we include~$\phi'(c) = v''$ itself. By Claim~\ref{claim:triangle:tightbudget} this shows that the subdivided star model~$\phi'$ does not leave enough activator vertices free to realize~$t$ disjoint triangles. On the other hand, if~$v \in V(\G)$ is used in the model~$\phi'$ of the subdivided star, still at least~$m$ activator vertices are used. But by Claim~\ref{claim:triangle:tightbudget}, if~$v \in V(\G)$ is not used to model a triangle, then we cannot pack~$t$ triangles using the~$|V(\G)| - m$ activator vertices that are left after discarding those used in the subdivided star. Again,~$\phi$ does not model~$H$ and we reach a contradiction. \end{itemize} As the vertices of~$G$ can be partitioned into instance selectors, dummies, activators, and vertices of~$V(\G)$, we find that if~$\phi$ is a full $H$-subgraph model in~$G$, then~$\phi(c)$ is an instance selector, say~$u_i$. As~$\deg_G(u_i) = \deg_H(c)$, all $G$-neighbors of~$u_i$ (which are activator vertices) are used to model vertices that subdivide the star. Hence there is a packing of~$t$ disjoint triangles in the graph~$G - N_G[u_i]$. Observe that for each vertex~$v \not \in S_i$, the graph~$G - N_G[u_i]$ contains~$v'$ and~$v''$. As these two vertices do not occur in any triangles except~$\{v,v',v''\}$, we may assume that the packing of triangles in~$G - N_G[u_i]$ contains all triangles~$\{v,v',v''\}$ for~$v \in V(\G) \setminus S_i$. This accounts for~$|V(\G)| - m$ of the~$t$ triangles. The remaining~$t - (|V(\G)| - m) = m/3$ triangles cannot use any dummy vertex, as dummy vertices only form triangles with activator vertices that have already been used fully. So the remaining~$m/3$ triangles are in fact triangles in~$G[S_i] = \G[S_i]$. As~$m = |S_i|$ this implies that~$\G[S_i] = \G^{K_3}_n[S_i]$ can be partitioned into triangles, which concludes the proof. \end{claimproof} \noindent As all criteria of an OR-cross-composition have been met, Theorem~\ref{theorem:karp:lowerbound:onesubstar:manytriangles} follows from Theorem~\ref{theorem:cross:composition:no:kernel}. \end{proof} \subsubsection{Two subdivided stars and many \texorpdfstring{$P_3$}{P3}'s} The next kernelization lower bound concerns pattern graphs that contain two subdivided stars together with many~$P_3$'s. The following lemma employs a padding argument to prove that \PathPackingCanonical remains NP-complete under some degree restrictions. These degree restrictions will be useful to during the OR-cross-composition to reason about how the center of a large subdivided star can appear in the constructed host graph. \begin{lemma} \label{lemma:canonicalpacking:larges:npcomplete} \PathPackingCanonical remains NP-complete when restricted to instances~$(n, S)$ where, if~$n \geq 10$, we have~$|S| > \Delta (\G^{P_3}_n) + 1$. \end{lemma} \begin{proof} Observe that for~$n \geq 10$ the maximum degree the graph~$\G^{P_3}_n$ is determined by the degree of the vertices~$U$ (which all have the same degree), since the private vertices of the gadgets have degree three. A vertex~$u \in U$ is adjacent in~$\G^{P_3}_n$ to one vertex of each gadget corresponding to a subset in~$\binom{U}{3}$ that contains~$u$. As there are~$\binom{|U| - 1}{2}$ such subsets, the degree of~$u \in U$ is~$\binom{n-1}{2}$. We can alter the NP-completeness transformation from \XTC to \PathPackingCanonical described in Lemma~\ref{lemma:canonicalpacking:npcomplete} as follows. Given an input~$(U, T)$ of \XTC with~$|U| = n$, add~$3n^5$ new elements~$x_1, y_1, z_1, \ldots, x_{n^5}, y_{n^5}, z_{n^5}$ to the universe to obtain~$U'$ and add all size-three subsets of these new elements to~$T$ to obtain~$T'$. It is easy to see that instance~$(U,T)$ is equivalent to instance~$(U',T')$. Now reduce~$(U',T')$ to an instance~$(n + 3 n^5, S)$ of \PathPackingCanonical as described in Lemma~\ref{lemma:canonicalpacking:npcomplete}. The maximum degree of the canonical graph~$\G^{P_3}_{n + 3 n^5}$ is at most~$\binom{n + 3 n^5 - 1}{2}$. Recall that~$S$ contains the vertices of the universe of the exact cover instance together with the gadgets created for triples that are contained in~$T'$. As~$T'$ contains at least~$\binom{3n^5}{3}$ sets, the constructed set~$S$ contains the nine private vertices of at least~$\binom{3n^5}{3}$ gadgets. As~$n \geq 10$, a simple computation shows that~$\Delta(\G^{P_3}_{n + 3 n^5}) \leq \binom{n + 3 n^5 - 1}{2} < 9 \binom{3n^5}{3} - 1 \leq |S| - 1$. As we pad with polynomially many new triples, the running time remains polynomial. Hence the overall construction of padding and then performing the construction of Lemma~\ref{lemma:canonicalpacking:npcomplete} gives a polynomial-time transformation from \XTC to the restricted form of \PathPackingCanonical, which concludes the proof. \end{proof} \restatekarptwostarmanyps* \begin{proof} Using the same polynomial equivalence relation as in Theorem~\ref{theorem:karp:lowerbound:onesubstar:manytriangles} and by discarding the same trivial cases, it suffices to OR-cross-compose instances~$(n, S_1), \ldots, (n, S_r)$ of \PathPackingCanonical with~$n \geq 10$ whose sets~$S_i$ all have the same size~$m \geq 10$. By Lemma~\ref{lemma:canonicalpacking:larges:npcomplete}, we may assume that~$m = |S_i| > \Delta (\G^{P_3}_n) + 1$ for all~$i \in [r]$. Let~$\G := \G^{P_3}_n$ be the canonical graph for these instances and define~$n' := |V(\G)|$. The construction of~$G$ is a bit more involved because we have to give a construction that embeds two subdivided stars, rather than one. We proceed as follows. Label the vertices of~$\G$ as~$v_1, \ldots, v_{n'}$ in an arbitrary way. \begin{itemize} \item Initialize~$G$ as a copy of the graph~$\G$. \item Add a matching~$M$ of size~$n'$ to the graph. For each index~$j \in [n']$ the $j$-th edge of~$M$ has the \emph{blocker vertex~$b_j$} as one endpoint and the \emph{propagator vertex~$p_j$} as the other endpoint. For each~$j \in [n']$, add the edge~$\{b_j, v_j\}$ to~$G$. \item Add another matching~$M'$ of size~$n'$ to the graph. For each index~$j \in [n']$ the $j$-th edge of~$M'$ has the \emph{dummy vertex~$d_j$} as one endpoint and the \emph{communicator vertex~$c_j$} as the other endpoint. For each~$j \in [n']$, add the edge~$\{c_j, p_j\}$ to~$G$. \item Add a special \emph{carving vertex}~$z^*$ and make it adjacent to the endpoints of matching~$M$. \item Add the \emph{carving matching}~$M''$ of size~$(n')^2$ to the graph and make one endpoint of each edge adjacent to~$z^*$. The carving matching ensures that there is a large subdivided star centered at the carving vertex, using~$M''$ and some other neighbors for $z^*$ for the legs of the star. This will force models of~$H$ to center one subdivided star at~$z^*$, as the constructed graph~$G$ will not contain other possibilities for the centers of such large subdivided stars. \item For each~$i \in [r]$, add an \emph{instance selector} vertex~$u_i$ to~$G$. For each vertex~$v_j \in S_i$ add the edge~$\{u_i, c_j\}$ to~$G$. \end{itemize} This concludes the description of~$G$. The graph~$H$ is defined as follows. It consists of~$m / 3$ copies of the graph~$P_3$, one subdivided star~$H_1$ with~$(2n' - m) + (n')^2$ leaves, and one subdivided star~$H_2$ with~$m$ leaves. Let~$y_1 \in V(H_1)$ be the center of~$H_1$ and let~$y_2 \in V(H_2)$ be the center of~$H_2$. From this choice it is clear that the parameter~$k$ of the constructed \kFSubgraphTest problem, which equals~$|V(H)|$, is polynomially bounded in the size of the largest input instance. It is easy to see that the construction can be performed in polynomial time. \begin{observation} \label{observation:twostars:lowerbound:firstcenter} The carving vertex~$z^*$ is the unique vertex of~$G$ for which both the first neighborhood~$N_G(z^*)$ and the second neighborhood~$N_G(N_G(z^*))$ have size at least~$(2n' - m) + (n')^2$: the vertices of~$V(\G)$ have degree in~$G$ at most~$\Delta(\G) + 1 \leq n' + 1 < (2n' - m) + (n')^2$, the vertices in~$V(M'')$, and the blocker, dummy, and propagator vertices have degree at most four, the instance selector vertices have degree~$m \leq n'$, and a communicator vertex~$c_j$ has second neighborhood~$N_G(N_G(c_j)) \subseteq \{b_j, d_j, p_j, z^*\} \cup \{c_\ell \mid \ell \in [n']\}$ of size at most~$4 + n' < (2n' - m) + (n')^2$. \end{observation} \begin{observation} \label{observation:twostars:lowerbound:secondcenter} As~$m > \Delta(\G) + 1 \geq \deg_G(v_j)$ for every $v_j \in V(\G)$, the only vertices in~$G$ that have degree at least~$m$ are the communicator vertices, the instance selector vertices, and~$z^*$. \end{observation} \begin{observation} \label{observation:twostars:wherearepaths} The only~$P_3$ subgraphs in~$(G - \{z^*\}) - \{b_j, c_j \mid j \in [n']\}$ are contained in~$G[V(\G)]$, as removing these vertices turns~$M''$ into an isolated matching while the dummy vertices, instance selector vertices, and propagator vertices become isolated. \end{observation} To establish the correctness of the OR-cross-composition we have to prove that~$H \subseteq G$ if and only if there is a \yes-instance among the inputs. To illustrate how the construction is intended to work, we first prove the reverse direction. \begin{claim}\label{claim:crosscomp:twostars:backwards} If~$i \in [r]$ such that~$\G[S_i]$ can be partitioned into~$P_3$'s, then~$H$ is a subgraph of~$G$. \end{claim} \begin{claimproof} Assume that~$\G[S_i]$ can be partitioned into~$P_3$'s. We construct a full $H$-subgraph model~$\phi$ in~$G$. As~$|S_i| = m$, a partition of~$\G[S_i]$ into~$P_3$'s is a packing of~$m/3$ vertex-disjoint~$P_3$'s. As~$\G$ is a subgraph of~$G$, there exists a packing of~$m/3$ vertex-disjoint~$P_3$'s in~$G[S_i]$. We show how to realize models of~$H_1$ and~$H_2$ that are disjoint from each other and from~$S_i$. The model of~$H_2$ is centered at~$u_i$. Recall that~$N_G(u_i) = \{c_j \mid v_j \in S_i\}$. These~$|S_i| = m$ vertices model the subdivider vertices of the subdivided star~$H_2$. We use the dummy vertices~$\{d_j \mid v_j \in S_i\}$ as the degree-one endpoints of the star~$H_2$. We obtain a realization of~$H_2$ using~$\{u_i\} \cup \{d_j, c_j \mid v_j \in S_i\}$. The model of~$H_1$ is centered at the carving vertex~$z^*$. Each of the~$(n')^2$ edges in the carving matching~$M''$ is used to realize one leg of the subdivided star. The remaining~$2n' - m$ legs are realized as follows. For each index~$j \in [n']$ such that~$v_j \in S_i$, we realize one leg of the subdivided star through the $j$-th edge~$\{p_j, b_j\}$ in the matching~$M$. For each~$j \in [n']$ such that~$v_j \not \in S_i$, we realize \emph{two} legs of the subdivided star: one leg~$\{p_j, c_j\}$ and one leg~$\{b_j, v_j\}$. From these definitions it is clear that the models of~$H_2$ and~$H_1$ are disjoint from each other and from~$S_i$. As we realize one leg of the subdivided star~$H_2$ for each of the~$m$ indices~$j$ with~$v_j \in S_i$, while we realize two legs for the~$n' - m$ indices~$j$ with~$v_j \not \in S_i$, together these realize all~$(n')^2 + m + 2(n' - m) = (2n' - m) + (n')^2$ legs of the subdivided star~$H_2$. Hence~$G$ contains~$H$ as a subgraph. \end{claimproof} The following series of claims will establish the reverse of Claim~\ref{claim:crosscomp:twostars:backwards}: if~$H \subseteq G$ then the answer to some input instance is \yes. \begin{claim} \label{claim:crosscomp:twostars:carvingcenter} If~$\phi$ is a full subgraph model of~$H$ in~$G$, then~$\phi(y_1) = z^*$. \end{claim} \begin{claimproof} If~$\phi$ is a full subgraph model of~$H$ in~$G$, then the degree of~$\phi(y_1)$ in~$G$ has to be at least~$\deg_H(y_1) = (2n' - m) + (n')^2$. By Observation~\ref{observation:twostars:lowerbound:firstcenter}, vertex~$z^*$ is the only possible choice. \end{claimproof} \begin{claim}\label{claim:crosscomp:twostars:manycommunicators} Let~$\phi$ be a full subgraph model of~$H$ in~$G$ and let~$C := \{c_j \mid j \in[n'] \wedge c_j \in \phi(V(H_1))\}$. The following hold. \begin{enumerate} \item $|C| \geq n' - m$. \item If~$|C| = n' - m$, then (a)~$\phi(V(H_1))$ contains all vertices~$\{v_j \mid c_j \in C\}$ and (b)~$\phi(V(H_1))$ contains all blocker vertices. \end{enumerate} \end{claim} \begin{claimproof} Let~$\phi$ be a full subgraph model of~$H$ in~$G$, which implies by the previous claim that~$\phi(y_1) = z^*$. Matching~$M''$ can realize~$(n')^2$ of the legs of~$H_1$ and, as~$\phi(y_1) = z^*$, it is easy to verify they cannot be useful in any other role. Consider how the remaining~$2n' - m$ legs of~$H_1$ are realized and observe that~$N_G(z^*) \setminus V(M'') = V(M)$. A priori, each vertex of~$N_G(z^*) \setminus V(M'')$ can be used to model a subdivider vertex of a leg of the star. However, for each~$j \in [n']$, if~$c_j \not \in \phi(V(H_1))$, then vertices~$\{p_j, b_j\}$ cannot both model subdivider vertex of different legs of the star: as~$N_G(p_j) = \{z^*, c_j, b_j\}$, the only way to utilize~$p_j$ to realize a leg without~$c_j$ is by using~$\{p_j, b_j\}$ as one leg, but then only one of the vertices of~$\{p_j, b_j\}$ models a subdivider vertex. It follows that if~$|C| < n' - m$, then there are more than~$n' - m$ edges~$\{p_j, b_j\}$ of~$M$ through which only one leg of the subdivided star is realized. Through the remaining edges of~$M$, at most two legs can be realized. Hence the subdivided star can realize less than~$(n')^2 + m + 2(n' - m)$, and consequently~$\phi$ does not model the subdivided star with~$(2n' - m) + (n')^2$ leaves. We conclude that~$|C| \geq n' - m$, proving the first part of the claim. For the proof of the second part, assume that~$|C| = n' - m$. Besides the~$(n')^2$ legs of~$H_1$ realized in~$M''$, there are~$(2n' - m) - |M'| = n' - m$ edges of~$M$ for which both endpoints of the edge realize subdivider vertices of different legs of the star. By the argument above, the vertices of an edge~$\{p_j, b_j\}$ of~$M$ can only model subdivider vertices of two different legs of the star if these legs are realized as~$\{p_j, c_j\}$ and~$\{b_j, v_j\}$. As each of these contributes one vertex to~$C$, while~$|C| = n' - m$, it follows that whenever~$c_j \in C$ the model must realize two legs of the star through~$\{p_j, b_j\}$ and therefore use~$\{b_j, v_j\}$ to model one leg of the star, forcing~$v_j \in \phi(V(H_1))$. This proves (a) of the second part of the claim. For~(b), observe that if~$b_j \not \in \phi(V(H_1))$, then either~$c_j \in \phi(V(H_1))$ while only one leg of the star is realized for the $j$-th edge of~$M$, or~$c_j \not \in \phi(V(H_1))$ and no legs of the star are realized for the $j$-th edge of~$M$. It follows that not all legs can be realized, a contradiction. \end{claimproof} \begin{claim} \label{claim:crosscomp:twostars:selectsinstance} If~$\phi$ is a full subgraph model of~$H$ in~$G$, then~$\phi(y_2) \in \{u_1, \ldots, u_r\}$. \end{claim} \begin{claimproof} Let~$\phi$ be a full subgraph model of~$H$ in~$G$. By Claim~\ref{claim:crosscomp:twostars:carvingcenter} we have~$\phi(y_1) = z^*$. By Observation~\ref{observation:twostars:lowerbound:secondcenter} and the fact that~$\deg_H(y_2) = m$, it follows that~$\phi(y_2)$ can only be an instance selector vertex or a communicator vertex. It remains to prove that the latter cannot happen. Suppose that~$\phi(y_2) = c_j$ for some~$j \in [n']$. Consider how the legs of the star~$H_2$ are realized in~$H$. Observe that~$N_G(c_j) \subseteq \{d_j, p_j\} \cup \{u_1, \ldots, u_r\}$. As~$\deg_G(d_j) = 1$ vertex~$d_j$ cannot form the subdivider vertex of a leg of the star. Hence~$p_j$ can be one subdivider vertex of the star~$H_2$ and the remaining subdivider vertices of the star~$H_2$ must be instance selectors. For each instance selector~$u_j$ that models a subdivider vertex, a neighbor~$N_G(u_j) \subseteq \{c_\ell \mid \ell \in [n']\}$ is used to model the degree-one endpoint of the corresponding leg of the star. We distinguish two cases. \begin{itemize} \item If~$p_j$ is not used to model a subdivider vertex of~$H_2$, then~$m$ instance selectors are used for this role, which leads to~$m$ communicator vertices other than~$\phi(y_2)$ being used in~$\phi(V(H_2))$. Including~$\phi(y_2)$ itself,~$\phi(V(H_2))$ contains~$m + 1$ communicator vertices. But then~$\phi(V(H_1))$, which is disjoint from~$\phi(V(H_2))$, realizes~$H_1$ using strictly less than~$n' - m - 1$ communicator vertices, a contradiction to Claim~\ref{claim:crosscomp:twostars:manycommunicators}. \item If~$p_j$ models a subdivider vertex of~$H_2$, then~$b_j$ is used for the degree-one endpoint of that leg of the star (as~$N_G(p_j) = \{z^*, c_j, b_j\}$ and~$z^* = \phi(y_1)$). Hence~$\phi(V(H_2))$ contains~$\{p_j, b_j\}$, the endpoints of the $j$-th edge of~$M$. The remaining~$m-1$ legs of the star are realized through instance selectors and their communicator neighbors, leading to an additional~$m-1$ communicator vertices being used in~$\phi(V(H_2))$. Hence~$\phi(V(H_2))$ contains~$m$ communicator vertices and the endpoints of one edge in~$M$. But then the model of~$H_1$, which is disjoint from~$\phi(V(H_2))$, contains at most~$n' - m$ communicator vertices and is disjoint from~$\{b_j, p_j\}$ for at least one~$j \in [n']$, contradicting Claim~\ref{claim:crosscomp:twostars:manycommunicators}. \end{itemize} This proves Claim~\ref{claim:crosscomp:twostars:selectsinstance}. \end{claimproof} \begin{claim} If~$\phi$ is a full subgraph model of~$H$ in~$G$ with~$\phi(y_1) = z^*$ and~$\phi(y_2) = u_i$ for some~$i \in [r]$, then~$\G[S_i]$ can be partitioned into~$P_3$'s. \end{claim} \begin{claimproof} Assume that the stated conditions hold. As~$N_G(u_i) = C_j := \{c_j \mid v_j \in S_i\}$ and the degree of~$u_i$ in~$G$ matches the degree of~$y_2$ in~$H_2$, it follows that all~$m$ vertices of~$C_j$ are used in~$\phi(V(H_2))$. Consider~$\overline{C_j} := \{c_j \mid j \in [n']\} \setminus C_j$. Since no vertex of~$C_j$ can be used in~$\phi(V(H_1))$, the model of~$H_1$ uses at most~$n' - m$ communicator vertices. By Claim~\ref{claim:crosscomp:twostars:manycommunicators} it follows that~$\phi(V(H_1))$ contains exactly~$n' - m$ communicator vertices, which must be those in~$\overline{C_j}$ since they are the only ones left. By the second part of Claim~\ref{claim:crosscomp:twostars:manycommunicators}, we find that~$\phi(V(H_1))$ contains~$\{v_j \mid c_j \in \overline{C_j}\}$ and all blocker vertices. Hence none of these vertices can be used to model~$P_3$'s. Similarly, no vertices of~$C_j \cup \overline{C_j}$ can be used to model a~$P_3$, nor can~$z^*$ be used since it is used as the center of~$H_1$. It follows that no vertices of~$\{z^*\} \cup \{c_j, b_j \mid j \in [n']\}$ are used to model~$P_3$'s, which implies by Observation~\ref{observation:twostars:wherearepaths} that all~$P_3$ models are contained in~$G[V(\G)]$. As they are disjoint from~$\phi(V(H_1))$, it follows that all~$P_3$ models are contained in~$G[V(\G)] - \{v_j \mid c_j \in \overline{C_j}\} = G[V(\G)] - \{v_j \mid v_j \not \in N_G(u_i)\} = G[V(\G)] - \{v_j \mid v_j \not \in S_i\} = G[S_i] = \G[S_i]$. Hence the~$m/3$ vertex-disjoint~$P_3$ models are contained in~$\G[S_i]$. As~$|S_i| = m$ this implies that~$\G[S_i]$ can be partitioned into~$P_3$'s. \end{claimproof} As the claims establish that instance~$(G,H)$ acts as the logical OR of the input instances, this concludes the proof of Theorem~\ref{theorem:karp:lowerbound:twosubstar:manyps}. \end{proof} \subsection{Upper bounds} \label{subsection:subgraph:karp:upperbounds} We now show that the kernelization lower bounds of Section~\ref{subsection:subgraph:karp:lowerbounds} are fragile in the sense that, if we slightly change the considered graph classes~$\F$, then the resulting \kFSubgraphTest problem admits polynomial (many-one) kernels. Both results we present in this section rely on the same underlying idea. The graph classes for which we obtained kernel lower bounds in Section~\ref{subsection:subgraph:karp:lowerbounds} are $(3,0,2,2)$-splittable. Theorem~\ref{theorem:kernel:subgraph} therefore gives Turing kernels for the \kFSubgraphTest problem on these graph families, which are based on guessing the model of a vertex set that realizes the split. In the many-one kernelization setting, this strategy fails since we have to produce a single, small output instance and therefore cannot cover the~$|V(G)|^c$ different options for the model of a vertex set that realizes the split. In the two cases highlighted below, the split is realized by just a single vertex and the mentioned problem can be circumvented: by ad-hoc arguments we can compute a representative vertex set~$Y$ of size~$k^{\Oh(1)}$ such that, if~$G$ contains~$H$ as a subgraph, then~$G$ has an $H$-subgraph model where the split vertex is realized by a member of~$Y$. Since the size of~$Y$ is polynomially bounded in the parameter, this allows us to compute representative sets for all relevant models of the split vertex. For each choice, Lemma~\ref{lemma:kernel:generic} gives a representative set of size polynomial in~$k$. The union of the representative sets over all~$y \in Y$ then gives us a kernel. \subsubsection{One subdivided star and many \texorpdfstring{$P_3$}{P3}'s} \label{section:subdividedstar:manypthrees} Throughout Section~\ref{section:subdividedstar:manypthrees}, we will use the term subdivided star for any graph that can be obtained from a star by subdividing each edge \emph{at most} once. Observe that under this definition, any path on at most three vertices is a subdivided star. If~$H$ is a subdivided star and~$v \in V(H)$ is the unique maximum-degree vertex in~$H$, then we will call~$v$ the \emph{center} of the subdivided star and we say that~$H$ has a center. A subgraph model of a subdivided star~$H$ in~$G$ is \emph{centered at~$v \in V(G)$} if~$v$ is the image of the center. In the degenerate case that~$H$ is a path on at most five vertices (which can be obtained by subdividing all edges of~$K_{1,2}$), all connected components of the pattern graph have constant size which allows us to obtain a polynomial kernel through Lemma~\ref{lemma:kernel:generic}. In the intermediate lemmas leading up to the kernelization, we therefore restrict ourselves to the cases that there is a unique center. We will need the following proposition for solving constrained weighted matching in polynomial time, which is possible by reducing it to an unconstrained minimum weight perfect matching computation. \begin{proposition}[{\cite{Plesnik99}}] \label{proposition:constrained:weighted:matching} There is a polynomial-time algorithm that, given an integer~$k$ and a graph~$G$ with nonnegative integer edge weights at most~$w_0$, computes in time polynomial in~$|V(G)| + |E(G)| + w_0$ a minimum-weight matching of cardinality~$k$, or determines that no matching of cardinality~$k$ exists. \end{proposition} Proposition~\ref{proposition:constrained:weighted:matching} allows us to find a subgraph model of a subdivided star in polynomial time, if one exists. While we could also perform this task using the randomized algorithm of Theorem~\ref{theorem:fsubgraphtest:alg}, the algorithm we present next has the advantage of being deterministic. \begin{lemma} \label{lemma:subdividedstartest:alg} There is a polynomial-time algorithm that, given a graph~$G$ and a subdivided star~$H$ with center~$c$ and a vertex~$v \in V(G)$, outputs a full $H$-subgraph model centered at~$v$ in~$G$ if one exists. \end{lemma} \begin{proof} Let~$k_1$ be the number of one-vertex components in~$H - \{c\}$ and let~$k_2$ be the number of two-vertex components in~$H - \{c\}$. Let~$G_c$ be the subgraph of~$G$ containing only the edges that have at least one endpoint in~$N_G(v)$. Define the weight of an edge in~$G_c$ as the number of endpoints the edge has in~$N_G(v)$; then every weight will be one or two. We invoke Proposition~\ref{proposition:constrained:weighted:matching} to compute a minimum-weight matching~$M_c$ in~$G_c$ of cardinality exactly~$k_2$. Let~$w_c$ be the weight of~$M_c$. \begin{claim} There is a full $H$-subgraph model~$\phi$ with~$\phi(c) = v$ if and only if the matching~$M_c$ exists and~$\deg_G(v) - w_c \geq k_1$. \end{claim} \begin{claimproof} For the forward direction, suppose that a matching of cardinality~$k_2$ exists in~$G_c$ and that~$M_c$ is a minimum-weight matching of this size of weight~$w_c$ satisfying~$\deg_G(v) - w_c \geq k_1$. We can realize a model of a subdivided star centered at~$v$ as follows: we use the edges in~$M_c$ for the two-vertex components of~$H - \{c\}$ (the legs of the star), which is possible since each edge has at least one endpoint that is adjacent to~$v$. By our choice of weight function we know that~$|N_G(v) \cap V(M_c)| = w_c$. By the degree requirement, we therefore find that~$N_G(v) \setminus V(M_c)$ consists of at least~$k_1$ vertices, which we can use to realize the size-one components of~$H - \{c\}$. Hence we obtain a full model of~$H$ in~$G$. For the reverse direction, suppose that~$\phi$ is a full $H$-subgraph model in~$G$ centered at~$v$. Since all the two-vertex components of~$H - \{v\}$ have at least one endpoint adjacent to~$v$, as this edge of~$H$ has to be realized, the model of the two-vertex components is a matching~$M$ in~$G_c$ of cardinality~$k_2$. The models of the one-vertex components consist of vertices of~$N_G(v) \setminus V(M)$, hence~$|N_G(v) \setminus V(M)| \geq k_1$. As~$M$ is a matching of cardinality~$k$ in~$G_c$, while~$M_c$ is a matching of cardinality~$k$ in~$G_c$ that minimizes the number of vertices of~$N_G(v)$ it uses (by our choice of weight function), it follows that~$|N_G(v) \setminus V(M)| \geq \deg_G(v) - w_c \geq k_1$, which concludes the reverse direction of the proof. \end{claimproof} The claim shows how to extract a model centered at~$v$ from the matching~$M_c$, if it exists. As~$M_c$ can be computed in polynomial time, the claim follows. \end{proof} \begin{lemma} \label{lemma:starsandpaths:findrelevantcenters} There is a polynomial-time algorithm with the following specifications. The input consists of a graph~$G$, a subdivided star~$H'$ with center~$c$, and a graph~$H''$ where each connected component is a path on at most three vertices. Let~$H := H' + H''$ and~$k := |V(H)|$. The output is a set~$Y \subseteq V(G)$ of size~$\Oh(k^2)$ such that if~$H \subseteq G$, then there is a full $H$-subgraph model in~$G$ in which the model of~$H'$ is centered at a vertex in~$Y$. \end{lemma} \begin{proof} On input~$(G,H',H'')$, the algorithm proceeds as follows. We first invoke Lemma~\ref{lemma:subdividedstartest:alg} to~$G$ and~$H'$ for all possible choices~$v \in V(G)$ for the center to determine whether~$G$ contains~$H'$ as a subgraph. If not, then there is no $H$-subgraph in~$G$ and we may output~$Y := \emptyset$. In the remainder we assume that~$H' \subseteq G$ and that~$\phi_{H'}$ is a full $H'$-subgraph model in~$G$. Let~$S$ be the set of vertices that have degree at least~$3k+1$ in~$G$. \begin{claim} If~$|S| \geq k$, then there is a full $H$-subgraph model in~$G$ where the model of~$H'$ is centered at~$\phi_{H'}(c)$. \end{claim} \begin{claimproof} Assume that~$|S| \geq k$. We construct an $H$-subgraph model~$\phi$ in~$G$, as follows. Define~$\phi(v) = \phi_{H'}(v)$ for all~$v \in V(H')$, which ensures that the center of the subdivided star is at~$\phi_{H'}(v)$. While there is a connected component~$C$ of~$H''$ that has not yet been assigned an image under~$\phi$, do the following. Let~$Z$ be the vertices in the current image of the partial $H$-subgraph model~$\phi$; then~$|Z| < k$ since the model is not yet complete. Hence there is at least one high-degree vertex~$s \in S \setminus Z$. As~$\deg_G(s) \geq 3k + 1$ and~$|Z| < k$, there are at least two vertices~$\{s_1, s_2\}$ in~$N_G(s) \setminus Z$. These three vertices form a path on three vertices in~$G$. As~$C$ is a path on at most three vertices, we can specify a model for~$C$ in~$\phi$ by mapping~$C$ to (a subset of)~$\{s, s_1, s_2\}$. By iterating this process we augment~$\phi$ to a full $H$-subgraph model in~$G$ with~$\phi(c) = \phi_{H'}(c) = v$, which proves the claim. \end{claimproof} The claim shows that if~$|S| \geq k$, then~$Y = \{\phi_{H'}(c)\}$ is a valid output for the procedure. In the remainder we therefore assume that~$|S| < k$. Let~$\P$ be a maximal packing of vertex-disjoint copies of~$P_3$ in the graph~$G-S$, which can be found by a greedy polynomial-time algorithm. \begin{claim} If~$\P$ contains at least~$k$ vertex-disjoint copies of~$P_3$, then there is a full $H$-subgraph model in~$G$ where the model of~$H'$ is centered at~$\phi_{H'}(c)$. \end{claim} \begin{claimproof} Assuming~$|\P| \geq k$ we show how to construct an $H$-subgraph model~$\phi$ in~$G$ with~$\phi(c) = \phi_{H'}(c)$. Define~$\phi(v) = \phi_{H'}(v)$ for all~$v \in V(H')$. While there is a connected component~$C$ of~$H''$ that has not yet been assigned an image under~$\phi$, let~$Z$ be the vertices currently used in the image of~$\phi$ and observe that~$|Z| < k$. Let~$P \in \P$ be a $P_3$-subgraph disjoint from~$Z$, which exists as~$|\P| \geq k$. As~$C$ is a subgraph of~$P$, we can map component~$C$ under~$\phi$ to (a subgraph of)~$P$. Iterating the procedure results in a full $H$-subgraph model as required. \end{claimproof} The claim shows that if~$|\P| \geq k$ then~$Y = \{\phi_{H'}(c)\}$ is a valid output for the procedure. In the remainder we assume that~$|\P| < k$. Let~$T$ be the vertices that occur in a~$P_3$ subgraph in~$\P$; then the previous assumption implies that~$|T| < 3k$. By our choice of~$\P$ as a maximal packing of~$P_3$'s, every connected component of~$G - (S \cup T)$ has at most two vertices. Let~$C_T$ be the connected components of~$G - (S \cup T)$ that are adjacent to a vertex in~$T$. As each vertex of~$T \subseteq V(G - S)$ has degree at most~$3k$, each vertex of~$T$ is adjacent to at most~$3k$ components in~$C_T$ and therefore~$|C_T| \leq 3k \cdot |T| \leq 9k^2$. Let~$T'$ be the union of~$T$ and the vertices in~$C_T$. As each component of~$C_T$ has at most two vertices,~$|T'| \leq |T| + 2 \cdot |C_T| \leq 3k + 18k^2$. By our choice of~$C_T$, the only vertices of~$G$ that a connected component of~$G - (S \cup T')$ can be adjacent to, are those in~$S$. Let~$X := V(G) \setminus (S \cup T')$. For each~$v \in X$ we invoke Lemma~\ref{lemma:subdividedstartest:alg} to determine whether there is an $H'$-subgraph model in~$G$ centered at~$v$. If no such subgraph model exists, then any $H$-subgraph in~$G$ must center the star at a vertex of~$V(G) \setminus X = S \cup T'$, so we may safely output~$Y := S \cup T'$ of size at most~$4k + 18k^2$. In the remainder of the proof we deal with the case that an $H'$-subgraph model can be centered in a vertex of~$X$. As the size of~$X$ is not bounded in~$k$, we cannot include all these vertices in the output set~$Y$. To deal with this issue, we will identify a single vertex~$x^* \in X$ such that if there is an $H$-subgraph model in~$G$ that centers the subdivided star at a member of~$X$, then there is such a model that centers the star at~$x^*$. The output set~$Y$ will then consist of~$S \cup T' \cup \{x^*\}$ of size at most~$4k + 18k^2 + 1$. It remains to define~$x^*$ and show that it has the claimed properties. We choose~$x^*$ as follows. \begin{itemize} \item If~$k_1 = 0$ then we let~$x^*$ be an arbitrary vertex of~$X$ such that~$G$ has an~$H'$-subgraph model centered there, which can be tested using Lemma~\ref{lemma:subdividedstartest:alg}. We say that we found an \emph{small} candidate for the center. \item If~$k_1 > 0$ then we have to be slightly more careful. Recall that all connected components of~$G[X] = G - (S \cup T')$ have at most two vertices. If there is a connected component~$C = \{u,v\}$ in~$G[X]$ such that~$G$ has an $H'$-subgraph model centered at~$u$, then we set~$x^* := u$ and we say that we found a \emph{large} candidate for the center. If no $H'$-subgraph model in~$G$ can be centered in a two-vertex component of~$G[X]$, then we let~$x^*$ be an arbitrary vertex of~$X$ that can be the center of an $H'$-subgraph model in~$G$ and we say that we found a \emph{small} candidate for the center. \end{itemize} Vertex~$x^*$ can be identified in polynomial time. The algorithm outputs the set~$Y := S \cup T' \cup \{x^*\}$. The following series of claims proves that this choice of~$Y$ satisfies the requirements in the statement of the claim. To simplify the notation in the rest of the proof, let~$\delta := 1$ if we found a small candidate for the center and~$0$ otherwise. \begin{claim} \label{claim:starsandpaths:model:uses:many:highdegree} Let~$\phi_{H'}$ be a full $H'$-subgraph model in~$G$ centered at a vertex~$x \in X$. Then $|\phi_{H'}(V(H')) \cap S| \geq k_1 + k_2 - \delta$. \end{claim} \begin{claimproof} We consider the three defining cases for~$x^*$ separately. \begin{itemize} \item If~$k_1 = 0$, then all connected components of~$H' - \{c\}$ consist of two vertices. Since~$\phi_{H'}(c)$ is contained in a connected component~$C$ of~$G[X]$, which has at most two vertices, no connected component~$C'$ of~$H' - \{c\}$ can have its image entirely within~$C$ because~$C$ also contains the image of~$c$. Hence the image of every such connected component~$C'$ of~$H' - \{c\}$ contains a vertex in~$N_G(C)$. By our choice of~$T'$, connected components of~$G[X] = G - (S \cup T')$ only have neighbors in~$S$. Hence the image of every connected component~$C'$ of~$H' - \{c\}$ intersects~$S$. As the images are vertex-disjoint and there are~$k_2$ such components, we find~$|\phi_{H'}(V(H')) \cap S| \geq k_2 = k_1 + k_2$ (as~$k_1 = 0$), which proves the claim in this case. \item If~$k_1 > 0$, and we found a large candidate for the center, then there is at most one connected component of~$H' - \{c\}$ whose image under~$\phi_{H'}$ does not intersect~$S$: if~$\phi_{H'}(c)$ is contained in a connected component~$C$ of~$G[X]$ of two vertices, then the vertex of~$C$ unequal to~$\phi_{H'}(c)$ can model a one-vertex component of~$H' - \{c\}$. The images of all other connected components intersect~$S$, by the same argument as in the previous case. Hence~$|\phi_{H'}(V(H')) \cap S| \geq k_1 + k_2 - 1$. \item If~$k_1 > 0$, but we found a small candidate for the center, then the mechanism for defining~$x^*$ ensures that it is impossible to have a~$H'$-subgraph model centered in a two-vertex component of~$G[X]$ where the non-center vertex of the component is used to model a one-vertex component of~$H' - \{c\}$. Hence all connected components of~$H' - \{c\}$ intersect~$S$, implying that~$|\phi_{H'}(V(H')) \cap S| \geq k_1 + k_2$. \end{itemize} This concludes the proof of Claim~\ref{claim:starsandpaths:model:uses:many:highdegree}. \end{claimproof} \begin{claim} \label{claim:starsandpaths:centeredmodel:fewhighdegree} For any set~$Z \subseteq T'$ of size at most~$k$ there is an $H'$-subgraph model~$\phi'$ in~$G - Z$ centered at~$x^*$ such that~$|\phi'(V(H')) \cap S| \leq k_1 + k_2 - \delta$. \end{claim} \begin{claimproof} We first prove the claim when we found a small candidate for the center. Afterwards we show how to adapt the proof in case of a large candidate for the center. \textbf{Small candidate.} Suppose we found a small candidate for the center. By the definition of~$x^*$, there is an $H'$-subgraph~$\phi'$ in~$G$ centered at~$x^*$. Recall that~$H - \{c\}$ has~$k_1 + k_2$ connected components. Each of the~$k_1$ one-vertex components can only contribute one to~$\phi'(V(H')) \cap S$. This implies that as long as $|\phi'(V(H')) \cap S| > k_1 + k_2$ holds, there is a two-vertex connected component~$C = \{u,v\}$ of~$H' - \{c\}$ with~$\{\phi'(u), \phi'(v)\} \subseteq S$. By the topology of a star, exactly one vertex of~$\{u,v\}$ is adjacent in~$H'$ to~$c$. Assume without loss of generality that this is~$v$, which implies that~$\{u,c\} \not \in E(H')$. Hence we may change the image of~$u$ without violating the validity of the $H'$-subgraph model, as long as the new image is disjoint from the rest of the model and is adjacent in~$G$ to~$\phi'(v)$. Since~$\phi'(v) \in S$ has degree more than~$2k$ in~$G$, there is a neighbor~$u'$ of~$\phi'(v)$ in~$G$ that is not used in the model~$\phi'$. By the preceding argument we may set~$\phi'(u) := u'$ to decrease the number of vertices of~$S$ used by the model by one, without violating the validity of the model. By iterating this argument we may assume that~$|\phi'(V(H')) \cap S| = k_1 + k_2$. We will now show how to alter~$\phi'$ to turn it into a model in~$G - Z$, i.e., how to avoid using vertices of~$Z$ in the image of~$\phi'$. Assume that~$z \in \phi'(V(H'))$. Since~$Z \subseteq T'$ and~$x^* \in X$ is not adjacent in~$G$ to any vertex in~$T'$, it follows that~$z$ is the image of a degree-1 endpoint~$e$ of the subdivided star that is not adjacent in~$H'$ to~$c$. Let~$d$ be the unique neighbor of~$e$ in~$H'$. Since~$\phi'(d)$ is adjacent to both~$x^* \in X$ and~$z \in T'$, we must have~$\phi'(d) \in S$ by our choice of~$T'$. Hence the degree of~$\phi'(d)$ in~$G$ is at least~$3k+1$. Since~$|Z| \leq k$, the current model uses at most~$k$ vertices, and~$|S| \leq k$, there is a vertex~$d'$ in~$N_G(\phi'(d)) \setminus (Z \cup \phi'(V(H')) \cup S)$. Consequently, we may update the model~$\phi'$ by setting~$\phi'(d) := d'$ while preserving a valid~$H'$-subgraph model without increasing the number of vertices of~$S$ used by the model. By iterating this procedure we arrive at a subgraph model of~$H'$ disjoint from~$Z$ (which is therefore a model in~$G - Z$) using exactly~$k_1 + k_2$ vertices in~$S$. \textbf{Large candidate.} The case where we found a large candidate for the center is similar; the main difference is that an $H'$-model can use one vertex of~$S$ less, by using the neighbor of~$x^*$ in~$G[X]$ as the image for a one-vertex component of~$H' - \{c\}$. In the case that we found a large candidate, the mechanism for defining~$x^*$ ensures that there is a~$H'$-subgraph~$\phi'$ in~$G$ centered at~$x^*$ such that~$x^*$ is contained in a two-vertex connected component~$C_{x^*} = \{x^*, y^*\}$ of~$G[X]$. We first show how to obtain a model~$\phi'$ where~$y^*$ is the image of a one-vertex component of~$H - \{c\}$. If~$y^* \not \in \phi'(V(H'))$, then we may take any one-vertex component of~$H - \{c\}$ (which exists since~$k_1 > 0$ if we found a large candidate) and map it to~$y^*$. Assume then that~$y^*$ is used as part of the image of a two-vertex component~$C_2$ of~$H - \{c\}$, and let~$C_1 = \{w\}$ be a one-vertex component of~$H - \{c\}$. Then~$w' := \phi'(w) \in S$ since~$N_G(x^*) \subseteq \{y^*\} \cup S$ while~$y^*$ is used for a different role in the model. Now we can swap the images of~$C_1$ and~$C_2$ under~$\phi$, as follows. We let~$\phi'(w) := y^*$. We map the vertex of~$C_2 \cap N_{H'}(c)$ to the vertex~$w' \in S$. Since~$w'$ has degree at least~$3k + 1$ in~$G$, it has a neighbor that is not yet used in the model; we use such a neighbor as the image of~$C_2 \setminus N_{H'}(c)$. We obtain a valid $H'$-subgraph model in~$G$ that is centered in~$x^*$ and in which the neighbor of~$x^*$ in~$G[X]$ is the image of a one-vertex component of~$H' - \{c\}$. While there is a connected component of~$H' - \{c\}$ whose image under~$\phi'$ contains two vertices in~$S$, we can update the model to reduce this number to one, just as in the previous case. Iterating this argument we obtain a model of~$H'$ centered at~$x^*$ where one component of~$H' - \{c\}$ does not use any vertex of~$S$, while the remaining components each use at most one vertex of~$S$. Hence this model~$\phi'$ satisfies~$|\phi'(H') \cap S| \leq k_1 + k_2 - 1$. Finally, just as in the previous case we may eliminate the vertices of~$Z$ from the model without increasing the number of~$S$-vertices that are used. This concludes the proof of Claim~\ref{claim:starsandpaths:centeredmodel:fewhighdegree}. \end{claimproof} \begin{claim} \label{claim:starsandpaths:centeratxprime} If there is a full $H$-subgraph model~$\phi$ in~$G$ in which the subdivided star~$H'$ is centered at a vertex~$x \in X$, then there is a full $H$-subgraph model in~$G$ in which the subdivided star~$H'$ is centered at~$x^*$. \end{claim} \begin{claimproof} We build a model~$\phi'$ of~$H$ in~$G$ that centers the subdivided star at~$x^*$. Let~$S' := \phi(V(H)) \cap S$. By Claim~\ref{claim:starsandpaths:model:uses:many:highdegree} we know that the~$H'$ submodel of~$\phi$ contributes at least~$k_1 + k_2 - \delta$ vertices to~$S'$. Hence there are at most~$|S'| - (k_1 + k_2 - \delta)$ vertices in~$S'$ that are the image of a vertex of a path in~$H$. Define~$\phi'$ as follows. \begin{itemize} \item For every path~$P$ in~$H''$ such that~$\phi(V(P)) \cap S = \emptyset$, set~$\phi'(v) = \phi(v)$ for all~$v \in V(P)$. Let~$Z \subseteq V(G)$ be the vertices used in the image of the partial model constructed in this way. \item By Claim~\ref{claim:starsandpaths:centeredmodel:fewhighdegree}, there is a $H'$-subgraph model~$\phi_{H'}$ of~$H'$ in~$G - Z$ that is centered at~$x^*$ and uses at most~$k_1 + k_2 - \delta$ vertices from~$S$. Set~$\phi'(v) = \phi_{H'}(v)$ for all~$v \in V(H')$; as~$\phi_{H'}$ is a model of~$G - Z$ no vertex is used twice in the partial model constructed so far. Let~$Z'$ be the vertices used by the model after this step. \item It remains to define an image for the connected components of~$H''$ whose image under~$\phi$ contains a vertex of~$S$. There are at most~$|S'| - (k_1 + k_2 - \delta) \leq |S| - (k_1 + k_2 - \delta)$ of such components by the observation above. On the other hand, there are at least~$|S| - (k_1 + k_2 - \delta)$ vertices in~$S$ that are not used by the partial model~$\phi'$ constructed so far. As each vertex of~$S$ has degree at least~$3k+1$ and the partial model contains at most~$k$ vertices, we can choose for each~$s \in S \setminus Z$ two vertices~$s_1, s_2 \in N_G(s) \setminus Z$ such that the assigned pairs are disjoint over all~$s \in S \setminus Z$. Each resulting triple is a path on three vertices in~$G$ that can form the image of one of the~$|S| - (k_1 + k_2 - \delta)$ remaining connected components of~$H''$. Hence there are sufficient triples to realize all remaining components of~$H''$. As~$H = H' + H''$ this gives a $H$-subgraph model in~$G$ centered at~$x^*$, as required. \end{itemize} \end{claimproof} As Claim~\ref{claim:starsandpaths:centeratxprime} shows that any model of~$H$ centered at a vertex of~$X$ can be transformed into a model centered at~$x^*$, the set~$Y := V(G) \setminus (X \setminus \{x^*\}) = S \cup T' \cup \{x^*\}$ satisfies the claimed requirements and may be used as the output. This concludes the proof of Lemma~\ref{lemma:starsandpaths:findrelevantcenters}. \end{proof} Using Lemma~\ref{lemma:starsandpaths:findrelevantcenters} we can prove the following theorem. \restatekarpstarspaths* \begin{proof} On input~$(G,H)$, the kernelization algorithm proceeds as follows. If~$H$ is not of the correct form, which is easily determined in polynomial time, then we output a constant-size \no-instance. If~$H'$ does not have a center then the subdivided star~$H'$ has at most five vertices, implying that~$H$ is $5$-small. Hence we obtain a polynomial kernel using Theorem~\ref{theorem:kernel:packing}. If~$H'$ has a unique center~$c$, then observe that all components of~$H'' := H - V(H')$ are paths on at most three vertices. We invoke Lemma~\ref{lemma:starsandpaths:findrelevantcenters} to compute a set~$Y$ of size~$\Oh(k^2)$ such that, if~$G$ contains~$H$ as a subgraph, then there is a subgraph model of~$H$ in~$G$ where~$c$ is mapped to a member of~$Y$. For each~$y \in Y$ construct a partial $H$-subgraph model~$\phi^{c \mapsto y}$ by setting~$\phi^{c \mapsto y}(c) := y$. Observe that every connected component of~$H - \{c\}$ has at most three vertices. Letting~$D := \{c\}$ the tuple~$(G,H,\phi^{c \mapsto y}, D)$ therefore satisfies the requirements of Lemma~\ref{lemma:kernel:generic} with~$(a,b,d) = (3, 0, 0)$. We may therefore invoke the lemma to compute a set~$X^{c \mapsto y}$ of size~$\Oh(k^{\Oh(a + b^2 + d)}) = \Oh(k^{\Oh(1)})$ such that if~$G$ has a full $H$-subgraph model extending~$\phi^{c \mapsto y}$, then~$G[X^{c\mapsto y}]$ contains a full $H$-subgraph model that extends~$\phi^{c \mapsto y}$. Let~$X$ be the union of~$X^{c \mapsto y}$ for all~$v \in Y$. As~$|Y| \in \Oh(k^2)$, the size of~$X$ is polynomial in~$k$. Lemma~\ref{lemma:starsandpaths:findrelevantcenters} guarantees that if~$G$ contains~$H$ as a subgraph, then there is a $H$-subgraph model in~$G$ that extends~$\phi^{c \mapsto y}$ for some~$v \in Y$ and therefore~$G[X]$ contains~$H$ as a subgraph. It follows that the instance~$(G[X], H)$ of \kFSubgraphTest, whose size is polynomial in~$k$, is equivalent to~$(G,H)$. As the set~$X$ can be computed in polynomial time, this gives a valid kernelization algorithm of polynomial size. \end{proof} \subsubsection{One fountain and many three-vertex components} In this section, we consider pattern graphs consisting of one (induced subgraph of a) fountain and many connected components of at most three vertices. It will be convenient to define a notion of center for a fountain, similarly as we did for subdivided stars. We will only concern ourselves with fountains whose cycle has length three; these are triangles where pendant vertices are attached to one vertex of the triangle. If such a fountain has a unique high-degree vertex, then this is the center of the fountain. In the degenerate case that there are no pendant vertices and the graph is two-regular, the graph has no center and the pattern graph~$H$ is $3$-small. As in the previous section we say that a subgraph model of a fountain~$H'$ in~$G$ is \emph{centered at~$v \in V(G)$} if the image of the center is~$v$. The main algorithmic tool will again be a lemma that identifies a small representative vertex set for the set of possible centers of $H$-subgraph models in~$G$. \begin{lemma} \label{lemma:fountainandtriangles:findrelevantcenters} There is a polynomial-time algorithm with the following specifications. The input consists of a graph~$G$, a graph~$H' \in \Ffountain{3}$ with center~$c$, and a graph~$H''$ where each connected component has at most three vertices. Let~$H := H' + H''$ and~$k := |V(H)|$. The output is a set~$Y \subseteq V(G)$ of size~$\Oh(k^3)$ such that if~$H \subseteq G$, then there is a full $H$-subgraph model in~$G$ in which the model of~$H'$ is centered at a vertex in~$Y$. \end{lemma} \begin{proof} Let~$c$ be the center of~$H'$ and let~$d,e$ be the two other vertices on the unique triangle in~$H'$. (Note that we require~$H' \in \Ffountain{3}$ which ensures that~$H'$ has a unique triangle; induced subgraphs of fountains that do not have a triangle do not satisfy the preconditions and will be dealt with separately.) We apply the sunflower lemma in a similar fashion as in Lemma~\ref{lemma:compute:representative:set:constanth} to compute the desired set~$Y$. Define a system of sets~$\S$ containing all triples~$\{c',d',e'\} \in \binom{V(G)}{3}$ for which there is a $H'$-subgraph model~$\phi$ in~$G$ with~$\phi(c) = c', \phi(d) = d'$ and~$\phi(e) = e'$. We can compute~$\S$ in polynomial time, since~$\{c',d',e'\} \in \binom{V(G)}{3}$ can realize~$H'$ if and only if~$\deg_G{c'} \geq \deg_H(c)$ and the triple induces a triangle in~$G$. We will compute a set~$\S' \subseteq \S$ with the following \emph{preservation property}: if~$G$ contains a full $H$-subgraph model, then~$G$ contains a full $H$-subgraph model~$\phi$ with~$\phi(\{c,d,e\}) \in \S'$. By our choice of~$\S$ it is clear that~$\S$ has the preservation property, so we initialize~$\S'$ as a copy of~$\S$. \begin{claim} \label{claim:fountainsandtriangles:preservationproperty} If~$\S' \subseteq \S$ has the preservation property and~$|\S'| \geq 6k^3$, we can identify a set~$S^* \in \S'$ in polynomial time such that~$\S' \setminus \{S^*\}$ also has the preservation property. \end{claim} \begin{claimproof} Suppose that~$|\S'| \geq 6k^3$. By Lemma~\ref{lemma:sunflowers}, there is a sunflower in~$\S'$ consisting of at least~$k+1$ sets~$S_1, \ldots, S_{k+1}$ and this can be found in time polynomial in the size of the set family and the universe. Let~$C := \bigcap _{i=1}^{k+1} S_i$ be the core of the sunflower. We show that~$\S' \setminus \{S_1\}$ has the preservation property. Suppose that~$\phi$ is a full $H$-subgraph model in~$G$. As~$\S'$ has the preservation property, there is a full $H$-subgraph model~$\phi'$ in~$G$ such that~$S_{\phi'} := \phi'(\{c,d,e\})$ is contained in~$\S'$. If~$S_{\phi'} \neq S_1$ then the set~$S_{\phi'}$ is contained in~$\S' \setminus \{S_1\}$, which establishes the preservation property. If~$S_1 = S_{\phi'}$, then we distinguish two cases depending on whether or not the core~$C$ is empty. \textbf{Empty core.} If~$C = \emptyset$, then sets~$S_1, \ldots, S_{k+1}$ induce vertex-disjoint triangles in~$G$ that each have a vertex of degree at least~$\deg_H(c)$. We show that there is an $H$-subgraph model~$\phi^*$ with~$\phi^*(c,d,e) = S_2$, thereby showing that~$S_1$ may be safely discarded. Let~$S_2 = \{c',d',e'\}$ and let~$c' \in V(G)$ satisfy~$\deg_G(c') \geq \deg_H(c)$. Define~$\phi^*(c) = c', \phi^*(d) = d'$, and~$\phi^*(e) = e'$. Let~$v_1, \ldots, v_{\deg_H(c)}$ be distinct vertices in~$N_G(c') \setminus \{c',d'\}$, which exist by the lower bound on the degree of~$c'$. We use~$v_1, \ldots, v_{\deg_H(c)}$ as the images under~$\phi^*$ of the pendant vertices attached to~$c$ in the fountain~$H'$. It remains to define images for the connected components in~$H''$. By the precondition to the lemma, each such component has at most three vertices. Observe that the set~$Z = \{c',d',e'\} \cup \{v_1, \ldots, v_{\deg_H(c)}\}$ has size at most~$|V(H')|$. As the sunflower has an empty core, the sets~$S_i$ are pairwise disjoint and therefore each vertex in~$Z$ intersects at most one set in the sunflower. Hence~$Z$ intersects less than~$|V(H')|$ sets in the sunflower. Consequently, there are at least~$k - |V(H')| = |V(H)| - |V(H')| = |V(H'')|$ sets among~$S_2, \ldots, S_{k+1}$ that are not intersected by~$Z$. As each set induces a triangle in~$G$, which is a supergraph of any connected component of~$H''$, each set not intersected by~$Z$ can form the image of one connected component of~$H''$. Hence there is a~$H$-subgraph model~$\phi^*$ in~$G$ with~$\phi^*(c,d,e) = S_2$, which shows that~$\S' \setminus \{S_1\}$ has the preservation property. \textbf{Non-empty core.} Now we deal with the case that~$C \neq \emptyset$. We construct a full $H$-subgraph model~$\phi^*$ in~$G$ such that~$\phi'(\{a,b,c\}) \in \{S_2, \ldots, S_{k+1}\}$. Recall that~$\phi'$ is the $H$-subgraph model we obtained from~$\S'$. For every vertex~$v \in V(H'')$, define~$\phi^*(v) = \phi'(v)$. Let~$Z := \bigcup _{v \in V(H'')} \phi'(v)$, which has size~$|V(H'')| < |V(H)| = k$. Hence there is a set~$S_i \in \{S_2, \ldots, S_{k + 1}\}$ that is disjoint from~$Z$. Let~$S_i = \{c_i, d_i, e_i\}$. As the core~$C$ is non-empty, at least one of the vertices in~$S_i$ is contained in~$C$; let~$c_i \in C$ and define~$\phi^*(c) = c_i, \phi^*(d) = d_i$, and~$\phi^*(e) = e_i$. It remains to define an image for the pendant vertices attached to~$c$ in the fountain. We claim that~$\deg_G(c_i) \geq k + 1$. To see this, observe that~$c_i$ is contained in the core of the sunflower, while each of the petals~$S_1 \setminus C, \ldots, S_{k+1} \setminus C$ contain at least one vertex that occurs in a common triangle with~$c_i$ (by definition of the sets in~$\S$). Hence each set in~$S_1 \setminus C, \ldots, S_{k+1} \setminus C$ contains a neighbor of~$c_i$, and as these sets are disjoint by the definition of a sunflower we indeed have~$\deg_G(c_i) \geq k+1$. The number of pendant vertices attached to~$c$ in~$H'$ is exactly~$\deg_{H}(c) = |V(H')| - 1$. As~$Z \cup \{b_i, c_i\}$ has size~$|V(H'')| + 2$, there are at least~$k+1 - (|V(H''| + 2) = |V(H')| - 1$ vertices in~$N_G(c_i) \setminus (Z \cup \{b_i, c_i\})$. Letting each such vertex form the image of one pendant neighbor of~$c$ in~$H'$, we can extend~$\phi^*$ to a full $H$-subgraph model. As~$\phi^*(\{c,d,e\}) = S_i \in \S' \setminus \{S_1\}$, the latter set has the preservation property and we can safely omit~$S_1$. This proves Claim~\ref{claim:fountainsandtriangles:preservationproperty}. \end{claimproof} By iterating the argument above, we arrive at a set system~$\S' \subseteq \S$ with the preservation property that contains at most~$6k^3$ sets. As each set in~$\S' \subseteq \S$ has size three, the set~$Y := \bigcup _{S \in \S'} S$ has size at most~$18k^3$. The preservation property directly implies that if there is a $H$-subgraph model in~$G$, then there is such a model that uses one of the triangles in~$\S'$ as the image for the triangle in the fountain, and that therefore centers the fountain at a member of~$Y$. Hence~$Y$ is a valid output for the procedure. As each iteration can be done in time polynomial in the size of~$\S$ and~$G$, while the number of iterations is bounded by~$|\S|$ which is~$\Oh(|V(G)|^3)$, the procedure runs in polynomial time. This concludes the proof of Lemma~\ref{lemma:fountainandtriangles:findrelevantcenters}. \end{proof} Lemma~\ref{lemma:starsandpaths:findrelevantcenters} allows us to derive a polynomial many-one kernel. \restatekarpfountaintriangles* \begin{proof} The main idea is the same as in the proof of Theorem~\ref{theorem:karp:subgraphkernel:starsandpaths}: find a representative set of centers of size polynomial in~$k$ and then invoke Theorem~\ref{lemma:kernel:generic} for each choice of center. As we are proving the theorem for the hereditary graph family~\F, we have to consider not just graphs~$H'$ that are contained in~$\Ffountain{3}$, but also their induced subgraphs. These are of two types: removing pendant vertices of a graph in \Ffountain{3} results in another graph in \Ffountain{3}. However, if we remove a vertex from the unique triangle of a graph in \Ffountain{3}, then this either reduces the graph to a star, or splits it into components with at most two vertices. Since Lemma~\ref{lemma:fountainandtriangles:findrelevantcenters} does not apply to stars, we have to deal with the case that~$H'$ is a member of \Ffountain{3} and that~$H'$ is a star separately. The kernelization algorithm works as follows. On input~$(G,H)$ it first tests whether~$H$ has the right form, which is easy to do in polynomial time. If this is not the case, it outputs a constant-size \no-instance. The behavior on the remaining instances depends on the form of~$H'$, which is a connected induced subgraph of a member of \Ffountain{3}. Recall that the parameter~$k$ is defined as~$|V(H)|$. If~$H' \in \Ffountain{3}$ and~$H'$ has a unique center~$c$ then we proceed as follows. We invoke Lemma~\ref{lemma:fountainandtriangles:findrelevantcenters} to compute a set~$Y \subseteq V(G)$ of size~$\Oh(k^3)$ such that, if~$G$ contains~$H$ as a subgraph, then there is a $H$-subgraph model that centers~$H'$ at a member of~$Y$. Let~$c$ be the center of~$H'$. For each~$y \in Y$ we create a partial $H$-subgraph model~$\phi^{c \mapsto y}$ with domain~$\{c\}$ that sets~$\phi^{c \mapsto y} := y$. Define~$D := \{c\}$. Just as in the proof of Theorem~\ref{theorem:karp:subgraphkernel:starsandpaths}, these parameters satisfy the requirements for Lemma~\ref{lemma:kernel:generic} with~$(a,b,d) = (3,0,0)$. We combine the resulting sets~$X^{c \mapsto y}$ over all choices of~$y$ into a set~$X$, and output the instance~$(G[X], H)$. Since~$|X| \in |Y| \cdot \Oh(k^{\Oh(a + b^2 + d)}) = \Oh(k^{\Oh(1)})$, the size of the reduced graph~$G[X]$ is bounded by a polynomial in~$k$. Correctness follows by exactly the same arguments as Theorem~\ref{theorem:karp:subgraphkernel:starsandpaths}. Hence we obtain a polynomial kernel if~$H' \in \Ffountain{3}$. If~$H' \not \in \Ffountain{3}$, then~$H'$ is a star (if~$H' \not \in \Ffountain{3}$) or~$H'$ is a triangle (if it has no center), which is a $1$-thin/$3$-small graph. Hence we obtain a polynomial kernel through Theorem~\ref{theorem:kernel:packing}. \end{proof} \bibliographystyle{abbrvurl}
\section{Introduction} \vspace{-0.35cm} In wireless communications, fading induced by multipath propagation has a detrimental effect on the received signals. Indeed, several fading can lead to a degradation of the transmission of information and the reliability of the network. In order to mitigate this effect, modern diversity techniques like cooperative diversity have been widely considered in recent years \cite{Proakis}. Several cooperative schemes have been proposed \cite{sendonaris,Venturino,laneman04} and among the most effective ones are Amplify-and-Forward (AF) and Decode-and-Forward (DF) \cite{laneman04,TDS_RS,TDS,smce,armo}. For an AF protocol, relays cooperate and amplify the received signals with a given transmit power, this method is simple except that the relays amplify their own noise. With DF protocol, relays decode the received signals and then forward the re-encoded message to the destination. Consequently, better performance and lower power consumption can be obtained when appropriate decoding and relay selection strategies are applied. DS-CDMA systems are a multiple access technique that can be incorporated with cooperative systems in ad hoc and sensor networks \cite{Bai,Souryal,Levorato,sit}. Due to the multiple access interference (MAI) effect that arises from nonorthogonal received waveforms, the system is adversely affected. To address this issue, multiuser detection (MUD) techniques have been developed in \cite{Verdu1} as an effective approach to suppress MAI. The optimal detector, known as maximum likelihood (ML) detector, has been proposed in \cite{Verdu2}. However, this method is infeasible for ad hoc and sensor networks considering its computational complexity. Motivated by this fact, several sub-optimal strategies have been developed: linear detectors \cite{Lupas}, successive interference cancellation (SIC) \cite{Patel,bsic,mbsic,Li1}, parallel interference cancellation (PIC) \cite{Varanasi,itic,mfpic} and decision feedback detectors \cite{mber,RCDL1,stmf,mfdf,dfjio,mbdf}. Prior studies on relay selection methods have been recently introduced in \cite{Jing,Clarke,Ding,Song,Talwar}. Among these approaches, a greedy algorithm is an effective way to approach the global optimal solution. Greedy algorithms have been widely applied in sparse approximation \cite{Tropp}, internet routing \cite{Flury} and arithmetic coding \cite{Jia}. In cooperative relaying systems, greedy algorithms are used in \cite{Ding,Song,lowbit} to search for the best relay combination. However, with insufficient number of combinations considered in the selection process, a significant performance loss is experience as compared to an exhaustive search. The objective of this paper is to propose a cross-layer design strategy that jointly considers the optimization of a low-complexity detection and a relay selection algorithm for ad hoc and sensor networks that employ DS-CDMA systems. Cross-layer designs that integrate different layers of the network have been employed in prior work \cite{RCDL2,Chen} to guarantee the quality of service and help increase the capacity, reliability and coverage of networks. However, there are very few works in the literature involving MUD techniques (operated in the lower physical layer) with relay selection (conducted in the data and link layer) in cooperative relaying systems. In \cite{Venturino,Cao}, an MMSE-MUD technique has been applied to cooperative systems, the results reveal that the transmissions become more resistant to MAI and obtain a significant performance gain when compared with a single direct transmission. However, extra complexity is introduced, as matrix inversions are required when an MMSE filter is employed. In this work, we devise a low-cost greedy list-based successive interference cancellation (GL-SIC) strategy with RAKE receivers as the front-end that can approach the maximum likelihood detector performance. Other more advanced front-ends \cite{mcg,cgb,ccmmwf,wlmwf,ifir,aifir1,aifir2,jio,stjio,smjio,jidf,sjidf,ccmrls,stbcccm,baidf} could be easily incorporated. Unlike prior art, the proposed GL-SIC algorithm exploits the Euclidean distance between users of interest and their nearest constellation points, with multiple ordering at each stage, we are able to build all possible lists of tentative decisions for each user via a greedy-like approach. We also present a low-cost multi-relay selection algorithm based on greedy techniques that can approach the performance of an exhaustive search. In the proposed greedy algorithm, a selection rule is employed and the process is separated into several stages. A cross-layer design technique that brings together the proposed GL-SIC algorithm and the greedy relay selection is then considered and evaluated by computer simulations. The rest of this paper is organized as follows. In Section 2, the system model is described. In Section 3, the GL-SIC multiuser detection method is presented. In Section 4, the relay selection strategy is proposed. In Section 5, simulation results are presented and discussed. Finally, conclusions are drawn in Section 6. \vspace{-0.6cm} \section{Cooperative DS-CDMA system model} \vspace{-0.35cm} We consider the uplink of a synchronous DS-CDMA system with $K$ users $(k_1,k_2,...k_K)$, $L$ relays $(l_1,l_2,...l_L)$, $N$ chips per symbol and $L_p$ $(L_p<N)$ propagation paths for each link. The system is equipped with a DF protocol at each relay and we assume that the transmit data are organized in packets comprising $P$ symbols. The received signals are filtered by a matched filter, sampled at chip rate to obtain sufficient statistics and organized into $M \times1$ vectors $\textbf{y}_{sd}$, $\textbf{y}_{sr}$ and $\textbf{y}_{rd}$, which represent the signals received from the sources (users) to the destination, the sources to the relays and the relays to the destination, respectively. The proposed algorithms for interference mitigation and relay selection are employed at the relays and at the destination. The received signal at the destination comprises the data transmitted during two phases that are jointly processed at the destination. Therefore, the received signal is described by a $2M\times1$ vector formed by stacking the received signals from the relays and the sources as given by \begin{equation} \begin{split} \hspace{-0.5em} \left[\hspace{-0.5em}\begin{array}{l} \textbf{y}_{sd}\\ \textbf{y}_{rd}\\ \end{array} \hspace{-0.5em} \right] & = \left[\hspace{-0.5em} \begin{array}{l} \sum\limits_{k=1}^K a_{sd}^k\textbf{S}_k\textbf{h}_{sd,k}b_k \\ \sum\limits_{l=1}^{L}\sum\limits_{k=1}^K a_{r_ld}^k\textbf{S}_k\textbf{h}_{r_ld,k}\hat{b}_{r_ld,k}\\ \end{array}\hspace{-0.5em} \right] + \left[ \hspace{-0.5em} \begin{array}{l} \textbf{n}_{sd}\\ \textbf{n}_{rd}\\ \end{array}\right], \label{equation1} \end{split} \end{equation} where $M=N+L_p-1$, $b_k\in\{+1,-1\}$ correspond to the transmitted symbols, $a_{sd}^k$ and $a_{r_ld}^k$ represent the $k$-th user's amplitude from the source to the destination and from the $l$-th relay to the destination. The $M \times L_p$ matrix $\textbf{S}_k$ contains the signature sequence of each user shifted down by one position at each column that forms \begin{equation} \textbf{S}_k = \left[\begin{array}{c c c } s_{k}(1) & & {\bf 0} \\ \vdots & \ddots & s_{k}(1) \\ s_{k}(N) & & \vdots \\ {\bf 0} & \ddots & s_{k}(N) \\ \end{array}\right], \end{equation} where $\textbf{s}_k=[s_k(1),s_k(2),...s_k(N)]^T$ is the signature sequence for user $k$. The vectors $\textbf{h}_{sd,k}$, $\textbf{h}_{r_ld,k}$ are the $L_p\times1$ channel vectors for user $k$ from the source to the destination and the $l$-th relay to the destination. The $M\times1$ noise vectors $\textbf{n}_{sd}$ and $\textbf{n}_{rd}$ contain samples of zero mean complex Gaussian noise with variance $\sigma^2$, $\hat{b}_{r_ld,k}$ is the decoded symbol for user $k$ at the output of relay $l$ after using the DF protocol. The received signal in (\ref{equation1}) can then be described by \begin{equation} \textbf{y}_d(i)=\sum\limits_{k=1}^K \textbf{C}_k \textbf{H}_k(i)\textbf{A}_k(i)\textbf{B}_k(i)+\textbf{n}(i), \end{equation} where $i$ denotes the time instant corresponding to one symbol in the transmitted packet and its received and relayed copies. $\textbf{C}_k$ is a $2M\times(L+1)L_p$ matrix comprising shifted versions of $\textbf{S}_k$ as given by \begin{equation} \textbf{C}_k = \left[\begin{array}{c c c c} \textbf{S}_{k} & {\bf 0} & \ldots & {\bf 0} \\ {\bf 0} & \ \ \textbf{S}_{k} & \ldots & \ \ \textbf{S}_{k}\\ \end{array}\right], \end{equation} $\textbf{H}_k(i)$ represents a $(L+1)L_p \times (L+1)$ channel matrix between the sources and the destination and the relays and the destination links. $\textbf{A}_k(i)$ is a $(L+1)\times(L+1)$ diagonal matrix of amplitudes for user $k$. The matrix $\textbf{B}_k(i)=[b_k,\hat{b}_{r_1d,k},\hat{b}_{r_2d,k},...\hat{b}_{r_Ld,k}]^T$ is a $(L+1)\times1$ vector for user $k$ that contains the transmitted symbol at the source and the detected symbols at the output of each relay, and $\textbf{n}(i)$ is a $2M\times1$ noise vector. \vspace{-0.35cm} \section{Proposed GL-SIC multiuser detection} \vspace{-0.35cm} In this section, we propose the GL-SIC multiuser detection method that can be applied at both the relays and the destination in the uplink of a cooperative system. The GL-SIC detector uses the RAKE receiver as the front-end, which reduces computational complexity by avoiding the matrix inversion required when MMSE filters are applied. The GL-SIC detector exploits the Euclidean distance between the users of interest and their nearest constellation points, with multiple ordering at each stage, we are able to build all possible lists of tentative decisions for each user. When seeking appropriate candidates, a greedy-like technique is performed to build each list and all possible estimates within the list are examined when unreliable users are detected. Unlike prior work which employs the concept of Euclidean distance with multiple feedback SIC (MF-SIC) \cite{Li1}, the proposed GL-SIC jointly considers multiple numbers of users, constellation constraints and re-ordering at each detection stage to obtain an improvement in detection performance. \vspace{-0.35cm} \subsection{The proposed GL-SIC design} \vspace{-0.35cm} \begin{figure}[!htb] \vspace{-0.7em} \begin{center} \def\epsfsize#1#2{1\columnwidth} \epsfbox{Fig2} \vspace{-2.5em} \caption{\footnotesize The reliability check for soft estimates in BPSK, QPSK and 16 QAM constellations.} \vspace{-1.5em} \label{fig2} \end{center} \end{figure} In the following, we describe the process for initially detecting $n$ users described by the indices $k_1, k_2,...,k_n$ at the first stage. Other users can be obtained accordingly. As shown by Fig. \ref{fig2}, $\beta$ is the distance between two nearest constellation points, $d_{th}$ is the threshold. The soft output of the RAKE receiver for user $k$ is then obtained by \begin{equation} u_{k}(i)=\textbf{w}_{k}^{H}\textbf{y}_{sr_l}(i), \end{equation} where $\textbf{y}_{sr_l}(i)$ represents the received signal from the source to the $l$-th relay, $u_{k}(i)$ stands for the soft output of the $i$-th symbol for user $k$ and $\textbf{w}_{k}^{H}$ denotes the RAKE receiver that corresponds to a filter matched to the effective signature at the receiver. After that, we order all users into a decreasing power level and organize them into a vector $\textbf{t}_a$. We pick the first $n$ entries $[\textbf{t}_a(1), \textbf{t}_a(2),...,\textbf{t}_a(n)]$ which denote users $k_1, k_2,...,k_n$, the reliability of each of the $n$ users is examined by the corresponding Euclidean distance between the desired user and its nearest constellation point $c$.\\\vspace{-1.2em} \\ \textbf{Decision reliable}:\\ If all $n$ users satisfy the following condition, they are determined as reliable, namely \begin{equation} u_{\textbf{t}_a(t)}(i)\notin \textbf{C}_{\rm grey},\ \ \ {\rm for} \ t\in [1,2,...,n], \end{equation} these soft estimates will be applied to a slicer $Q(\cdot)$ as described by \begin{equation} \hat{b}_{\textbf{t}_a(t)}(i)=Q(u_{\textbf{t}_a(t)}(i)),\ {\rm for}\ t\in [1,2,...,n], \end{equation} where $\hat{b}_{\textbf{t}_a(t)}(i)$ denotes the detected symbol for the $\textbf{t}_a(t)$-th user, $\textbf{C}_{\rm grey}$ is the shadowed area in Fig. \ref{fig2}, it should be noted that the shadowed region would spread along both the vertical and horizontal directions. After that, the cancellation is then performed in the same way as a conventional SIC where we mitigate the impact of MAI brought by these users: \begin{equation} \textbf{y}_{sr_l,s+1}(i)=\textbf{y}_{sr_l,s}(i)-\sum\limits_{t=1}^{n}\textbf{H}_{sr_l,\textbf{t}_a(t)}(i) \hat{b}_{\textbf{t}_a(t)}(i), \label{equation2} \end{equation} where $\textbf{H}_{sr_l,\textbf{t}_a(t)}(i) = a_{sr_l}^{\textbf{t}_a(t)}\textbf{S}_{\textbf{t}_a(t)}(i)\textbf{h}_{sr_l,\textbf{t}_a(t)}(i)$ stands for the desired user's channel matrix associated with the link between the source and the $l$-th relay, $\textbf{y}_{sr_l,s}$ is the received signal from the source to the $l$-th relay at the $s$-th $(s=1,2,...,K/n)$ cancellation stage. The above process is repeated subsequently with another $n$ users being selected from the remaining users at each following stage, and this algorithm changes to the unreliable mode when unreliable users are detected. Additionally, since the interference created by the previous users with the strongest power has been mitigated, improved estimates are obtained by reordering the remaining users.\\\vspace{-1.2em} \\ \textbf{Decision unreliable}:\\ (a). If part of the $n$ users are determined as reliable, while others are considered as unreliable, we have \begin{equation} u_{\textbf{t}_p(t)}(i) \notin \textbf{C}_{\rm grey},\ \ \ {\rm for}\ t\in [1,2,...,n_p], \label{equation3} \end{equation} \begin{equation} u_{\textbf{t}_q(t)}(i) \in \textbf{C}_{\rm grey},\ \ \ {\rm for}\ t\in [1,2,...,n_q], \label{equation4}\vspace{-0.8em} \end{equation} where $\textbf{t}_p$ is a $1 \times n_p$ vector that contains $n_p$ reliable users and $\textbf{t}_q$ is a $1 \times n_q$ vector that includes $n_q$ unreliable users, subject to $\textbf{t}_p\cap \textbf{t}_q=\varnothing$ and $ \textbf{t}_p\cup \textbf{t}_q=[1,2,...n]$ with $n_p+n_q=n$. Consequently, the $n_p$ reliable users are applied to the slicer $Q(\cdot)$ directly and the $n_q$ unreliable ones are examined in terms of all possible constellation values $c^m$ $(m=1,2,...,N_c)$ from the $1\times N_c$ constellation points set $\textbf{C}\subseteq \textsl{F}$, where $\textsl{F}$ is a subset of the complex field and $N_c$ is determined by the modulation type. The detected symbols are given by \begin{equation} \hat{b}_{\textbf{t}_p(t)}(i)=Q(u_{\textbf{t}_p(t)}(i)), {\rm for}\ t\in [1,2,...,n_p], \end{equation} \begin{equation} \hat{b}_{\textbf{t}_q(t)}(i)=c^m,\ \ {\rm for}\ t\in [1,2,...,n_q],\vspace{-0.7em} \end{equation} At this point, $N_c^{n_q}$ combinations of candidates for $n_q$ users are generated. The detection tree is then split into $N_c^{n_q}$ branches. After this processing, (\ref{equation2}) is applied with its corresponding combination to ensure the influence brought by the $n$ users is mitigated. Following that, $N_c^{n_q}$ updated $\textbf{y}_{sr_l}(i)$ are generated, we reorder the remaining users at each cancellation stage and compute a conventional SIC with RAKE receivers for each branch. The following $K\times1$ different ordered candidate detection lists are then produced \begin{equation} \textbf{b}^j(i)=[\textbf{s}_{\textrm{pre}}(i), \ \ \textbf{s}_p(i),\ \ \textbf{s}^j_q(i),\ \ \textbf{s}^j_{\textrm{next}}(i)]^T, \ j=1,2,...,N_c^{n_q}, \end{equation} where $\textbf{s}_{\textrm{pre}}(i)=[\hat{b}_{\textbf{t}_a(1)}(i),\hat{b}_{\textbf{t}_a(2)}(i),...]^T$ stands for the previous stages detected reliable symbols, $\textbf{s}_p(i)=[\hat{b}_{\textbf{t}_p(1)}(i), \hat{b}_{\textbf{t}_p(2)}(i)\\,...,\hat{b}_{\textbf{t}_p(n_p)}(i)]^T$ is a $n_p\times 1$ vector that denotes the current stage reliable symbols detected directly from slicer $Q(\cdot)$ when (\ref{equation3}) occurs, $\textbf{s}^j_q(i)=[c_{\textbf{t}_q(1)}^{m},c_{\textbf{t}_q(2)}^{m},...,c_{\textbf{t}_q(n_q)}^{m}]^T, j=1,2,...,N_c^{n_q}$ is a $n_q\times 1$ vector that contains the detected symbols deemed unreliable at the current stage under the situation of (\ref{equation4}), each entry of this vector is selected randomly from the constellation point set $\textbf{C}$ and all possible $N_c^{n_q}$ combinations need to be considered and examined. $\textbf{s}^j_{\textrm{next}}(i)=[...,\hat{b}_{\textbf{t}'(1)}^{\textbf{s}^j_q}(i),...,\hat{b}_{\textbf{t}'(n)}^{\textbf{s}^j_q}(i)]^T$ includes the corresponding detected symbols in the following stages after the $j$-th combination of $\textbf{s}_q(i)$ is allocated to the unreliable user vector $\textbf{t}_q$, and $\textbf{t}'$ is a $n \times 1$ vector that contains the users from the last stage.\\\vspace{-1.2em} \\ (b). If all $n$ users are considered as unreliable: \begin{equation} u_{\textbf{t}_b(t)}(i) \in \textbf{C}_{\rm grey},\ \ \ {\rm for}\ t\in [1,2,...,n],\label{equation5} \end{equation} where $\textbf{t}_b=[1,2,...,n]$, then all $n$ unreliable users can assume the values in $\textbf{C}$. In this case, the detection tree will be split into $N_c^{n}$ branches to produce \begin{equation} \hat{b}_{\textbf{t}_b(t)}(i)=c^m, \ {\rm for}\ t\in [1,2,...,n], \end{equation} Similarly, (\ref{equation2}) is then applied and a conventional SIC with different orderings at each cancellation stage is performed via each branch. Since all possible constellation values are tested for all unreliable users, we have the candidate lists: \begin{equation} \textbf{b}^j(i)=[\textbf{s}_{\textrm{pre}}(i),\ \textbf{s}_b^j(i),\ \ \textbf{s}^j_{\textrm{next}}(i)]^T, \ j=1,2,...,N_c^n, \end{equation} where $\textbf{s}_{\textrm{pre}}(i)=[\hat{b}_{\textbf{t}_a(1)}(i),\hat{b}_{\textbf{t}_a(2)}(i),...]^T$ are the reliable symbols that are detected from previous stages, $\textbf{s}_b^j(i)=[c_{\textbf{t}_b(1)}^{m},c_{\textbf{t}_b(2)}^{m},...,c_{\textbf{t}_b(n)}^{m}]^T, j=1,2,...,N_c^n$ is a $n\times1$ vector that represents $n$ number of users which are regarded as unreliable at the current stage as shown by (\ref{equation5}), each entry of $\textbf{s}_b^j$ is selected randomly from the constellation point set $\textbf{C}$. The vector $\textbf{s}^j_{\textrm{next}}(i)=[...,\hat{b}_{\textbf{t}'(1)}^{\textbf{s}_b^j}(i),...,\hat{b}_{\textbf{t}'(n)}^{\textbf{s}_b^j}(i)]^T$ contains the corresponding detected symbols in the following stages after the $j$-th combination of $\textbf{s}_b(i)$ is allocated to all unreliable users. After the candidates are generated, lists are built for each group of users, and the ML rule is used at the end of the detection procedure to choose the best candidate list as described by \begin{equation} \textbf{b}^{\textrm{best}}(i)= \min _{\substack{1\leq j\leq m, \textrm{where}\\m=N_c^{n_q} \textrm{or}\ N_c^n}}\parallel \textbf{y}_{sr_l}(i)-\textbf{H}_{sr_l}(i)\textbf{b}^{j}(i)\parallel^2. \end{equation} \vspace{-0.55cm} \subsection{GL-SIC with multi-branch processing} \vspace{-0.35cm} The multiple branch (MB) structure employs multiple parallel processing branches which can help to regenerate extra detection diversity has been discussed carefully in \cite{RCDL1,Li2}. Inspired by that, an extended structure that incorporates the proposed GL-SIC with multi-branch is achieved. By changing the obtained best detection order for $\textbf{b}^{\textrm{best}}$ with indices $\textbf{O}=[1,2,...,K]$ into a group of different detection sequences, we are able to gain this parallel structure with each branch shares a different detection order. Since it is not practical to test all $L_b=K!$ possibilities due to the high complexity, a reduced number of branches can be tested, with each index number in $\textbf{O}_{l_b}$ being the corresponding index number in $\textbf{O}$ cyclically shifted to right by one position, namely: $\textbf{O}_{l_1}=[K,1,2,...,K-2,K-1], \textbf{O}_{l_2}=[K-1,K,1,...,K-3,K-2],..., \textbf{O}_{l_{K-1}}=[2,3,4,...,K,1]$ and $\textbf{O}_{l_K}=[K,K-1,...,1]$(reversed order). After that, each of the $K$ parallel branches computes a GL-SIC algorithm with its corresponding order, respectively. After obtaining $K+1$ ($\textbf{O}$ included) different candidate lists according to each of the branch, a modified ML rule is applied with the following steps: \begin{enumerate} \item Obtain the best detection candidate branch $\textbf{b}^{O_{l_{\rm base}}}(i)$ among all $K+1$ parallel branches according to the ML rule \begin{equation} \textbf{b}^{O_{l_{\rm base}}}(i)= \min _{\substack{0\leq b\leq K}}\parallel \textbf{y}_{sr_l}(i)-\textbf{H}_{sr_l}\textbf{b}^{O_{l_b}}(i)\parallel^2.\label{equation6} \end{equation} \item Re-examine the detected value for user $k$ $(k=1,2,...,K)$ by fixing the detected results of all other unexamined users in $\textbf{b}^{O_{l_{\rm base}}}(i)$, then replace the $k$-th user's detection result $\hat b_{k}$ in $\textbf{b}^{O_{l_{\rm base}}}(i)$ by its corresponding detected values from all other $K$ branches $\textbf{b}^{O_{l_b}}(i)$ $(b=0,1,...,K, l_b\neq l_{\rm base}, \textbf{O}=\textbf{O}_{l_0})$ with the same index, the combination with the minimum Euclidean distance is selected through (\ref{equation6}) and the corresponding improved estimate of user $k$ is saved and kept. \item The same process is then automatically repeated with the next user in $\textbf{b}^{O_{l_{\rm base}}}(i)$ until all users in $\textbf{b}^{O_{l_{\rm base}}}(i)$ are examined. \end{enumerate} As each combination selection seeks an improvement based on the previous choice, a better candidate list $\textbf{b}^{\textrm{opt}}(i)$ can be generated. This is performed at all the relays and the destination using the RAKE receiver structure. \vspace{-0.45cm} \section{proposed greedy multi-relay selection method} \vspace{-0.35cm} In this section, a greedy multi-relay selection method is introduced. For this problem, the best relay combination is obtained through an exhaustive search of all possible subsets of relays. However, the major problem that prevents us from applying an exhaustive search in practical communications is its high computational complexity. With $L$ relays involved in the transmission, an exponential complexity of $2^L-1$ is experienced. This fact motivates us to seek other alternative methods. The standard greedy algorithm can be used in the selection process by cancelling the poorest relay-destination link stage by stage, however this method can approach only a local optimum. The proposed greedy multi-relay selection method can go through a reduced number of relay combinations and approach the best one based on previous decisions. In the proposed relay selection, the signal to interference and noise ratio (SINR) is used as the criterion to determine the optimum relay set. The expression of the SINR is given by \begin{equation} {\rm SINR_q} =\frac{E[|\textbf{w}_q^H\textbf{r}|^2]}{E[|\boldsymbol\eta|^2]+\textbf{n}}, \end{equation} where $\textbf{w}_q$ denotes the RAKE receiver for user $q$, $E[|\boldsymbol\eta|^2]=E[|\sum\limits_{\substack{k=1\\k\neq q}}^K\textbf{H}_kb_k|^2]$ is the interference brought by all other users, $\textbf{n}$ is the noise. For the RAKE receiver, the SINR is derived by \begin{equation} {\rm SINR_q} =\frac{\textbf{h}_q^H\textbf{H}\textbf{H}^H\textbf{h}_q} {\textmd{trace}(\textbf{H}_\eta \textbf{H}_\eta^H)+\textbf{h}_q^H\sigma_N^2\textbf{h}_q}, \end{equation} where $\textbf{h}_q$ is the channel vector for user $q$, $\textbf{H}$ is the channel matrix for all users, $\textbf{H}_\eta$ represents the channel matrix of all other users except user $q$. It should be mentioned that in various relay combinations, the channel vector $\textbf{h}_q$ for user $q$ $(q=1,2,...,K)$ is different as different relay-destination links are involved, $\sigma_N^2$ is the noise variance. This problem thus can be cast as the following optimization: \begin{equation} {\rm SINR_{\Omega_{best}}}= max \ \ \{min({\rm SINR_{\Omega_{r(q)}}}), q=1,...,K\}, \end{equation} where $\Omega_{r}$ denotes all possible combination sets $(r \leq L(L+1)/2)$ of any number of selected relays, ${\rm SINR_{\Omega_{r(q)}}}$ represents the SINR for user $q$ in set $\Omega_r$, min $({\rm SINR_{\Omega_{r(q)}}}) = {\rm SINR_{\Omega_r}}$ stands for the SINR for relay set $\Omega_r$ and $\Omega_{\rm best}$ is the best relay set that provides the highest SINR. \vspace{-0.5cm} \subsection{Standard greedy relay selection algorithm} \vspace{-0.35cm} The standard greedy relay selection method is operated in stages by cancelling the single relay according to the channel condition, as the channel path power is given by \begin{equation} P_{h_{r_ld}}=\textbf{h}_{r_ld}^H\textbf{h}_{r_ld}, \end{equation} where $\textbf{h}_{r_ld}$ is the channel vector between the $l$-th relay and the destination. The selection begins with all $L$ relays participating in the transmission, and the initial SINR is determined when all relays are involved in the transmission. For the second stage, we cancel the poorest relay-destination link and calculate the current SINR for the remaining $L-1$ relays, as compared with the previous SINR, if \begin{equation} {\rm SINR_{cur}}>{\rm SINR_{pre}}, \label{equation7} \end{equation} we update the previous SINR as \begin{equation} {\rm SINR_{pre}} = {\rm SINR_{cur}}, \label{equation8} \end{equation} and move to the third stage, where we remove the current poorest link and repeat the above process. The algorithm will stop when ${\rm SINR_{cur}}<{\rm SINR_{pre}}$ or when there is only one relay left. The whole process can be performed once before each packet transmission. \vspace{-0.5cm} \subsection{Proposed greedy relay selection algorithm} \vspace{-0.35cm} In order to improve the performance of the existing algorithm, we have modified the above standard greedy algorithm. The proposed method differs from the standard technique as we drop each of the relays in turns rather than drop them based on the channel condition at each stage. The algorithm can be summarized as: \begin{enumerate} \item Initially, a set $\Omega_A$ that includes all $L$ relays is generated and its corresponding SINR is calculated, denoted by ${\rm SINR_{pre}}$. \item For the second stage, we calculate the SINR for $L$ combination sets with each dropping one of the relays from $\Omega_A$. After that, we pick the combination set with the highest SINR for this stage, recorded as ${\rm SINR_{cur}}$. \item Compare ${\rm SINR_{cur}}$ with the previous stage ${\rm SINR_{pre}}$, if (\ref{equation7}) is true, we save this corresponding relay combination as $\Omega_{\textrm{cur}}$ at this stage. Meanwhile, we update the ${\rm SINR_{pre}}$ as in (\ref{equation8}). \item After moving to the third stage, we drop relays in turn again from $\Omega_{\textrm{cur}}$ obtained in stage two. $L-1$ new combination sets are generated, we then select the set with the highest SINR and repeat the above process in the following stages until either ${\rm SINR_{cur}}<{\rm SINR_{pre}}$ or there is only one relay left. \end{enumerate} This new greedy selection method considers the combination effect of the channel condition, which implies additional useful sets are examined. When compared with the standard greedy relay selection method, the previous stage decision is more accurate and the global optimum can be approached more closely. Similarly, the whole process is performed only once before each packet. Meanwhile, its complexity is less than $L(L+1)/2$, much lower than the exhaustive search.\vspace{-1.2em} \begin{table}[!htp]\scriptsize \footnotesize \centering\caption{The proposed greedy multi-relay selection algorithm} \begin{tabular}{l} \hline $\Omega_A=[1,2,3,...L]$\% $\Omega_A$ denotes the set when all relays are involved\\ ${\rm SINR_{\Omega_A}}=\textrm{min}({\rm SINR_{\Omega_{A(q)}}}), q=1,2,...K$\\ ${\rm SINR_{pre}}={\rm SINR_{\Omega_A}}$\\ \textbf{for} stage =1 \textbf{to} $L-1$\\ \ \ \ \ \ \textbf{for} $r$=1 \textbf{to} $L+1$-stage\\ \ \ \ \ \ \ \ \ \ $\Omega_r=\Omega_A-\Omega_{A(r)}$\% drop each of the relays in turns\\ \ \ \ \ \ \ \ \ \ ${\rm SINR_{\Omega_r}}=\textrm{min}({\rm SINR_{\Omega_r(q)}}), q=1,2,...,K$\\\ \ \ \ \ \textbf{end for}\\ \ \ \ \ \ ${\rm SINR_{\textrm{cur}}}=\textrm{max}({\rm SINR_{\Omega_r}})$\\ \ \ \ \ \ $\Omega_{\textrm{cur}}=\Omega_{{\rm SINR_{cur}}}$\\ \ \ \ \ \ \textbf{if} ${\rm SINR_{cur}}>{\rm SINR_{pre}}$ \textbf{and} $\rm{length}(\Omega_{\textrm{cur}})>1$\\ \ \ \ \ \ \ \ \ \ $\Omega_A=\Omega_{\textrm{cur}}$\\ \ \ \ \ \ \ \ \ ${\rm SINR_{pre}}={\rm SINR_{cur}}$\\ \ \ \ \ \ \textbf{else}\\ \ \ \ \ \ \ \ \ \ \ \textbf{break}\\ \ \ \ \ \textbf{end if}\\ \textbf{end for}\\ \hline \end{tabular} \vspace{-1.2em} \end{table} \vspace{-0.25cm} \section{simulations} \vspace{-0.35cm} In this section, a simulation study of the proposed GL-SIC multiuser detection strategy with a RAKE receiver and the low cost greedy multi-relay selection method is carried out. The DS-CDMA network uses randomly generated spreading codes of length $N=16$ and $N=32$, it also employs $L_p=3$ independent paths with the power profile $[0\rm dB,-3\rm dB,-6\rm dB]$ for each source-relay, source-destination and relay-destination link, their corresponding channel coefficients $h_{sr_l}^{l_p}$, $h_{sd}^{l_p}$ and $h_{r_ld}^{l_p} (l_p=1,2,...,L_p)$ are taken as uniformly random variables and normalized to ensure the total power is unity. We assume perfectly known channels at the receiver. Equal power allocation with normalization is assumed to guarantee no extra power is introduced during the transmission. We consider packets with 1000 BPSK symbols and average the curves over 300 trials. For the purpose of simplicity, $d_{th}=0.25$ and BPSK modulation technique are applied in the following simulations. Furthermore, in the GL-SIC algorithm, $n=2$ users are considered at each stage. The performance is evaluated in both non-cooperative and cooperative schemes. \vspace{-0.3cm} \begin{figure}[!htb] \begin{center} \def\epsfsize#1#2{0.8\columnwidth} \epsfbox{Fig3} \vspace{-1.5em}\caption{\footnotesize Non-cooperative system with N=32, 20 users over Rayleigh fading channel} \label{fig3} \end{center} \vspace{-1.5em} \end{figure} The first example shown in Fig. \ref{fig3} is the proposed GL-SIC interference suppression technique with 20 users that only takes into account the source-destination link. The conventional SIC detector is the standard SIC with RAKE receivers employed at each stage and the Multi-branch Multi-feedback (MB-MF) SIC detection algorithm mentioned in \cite{Li1} is presented here for comparison purposes. We also produced the simulation results for the multi-branch SIC (MB-SIC) where four parallel branches ($L_b=4$) with different detection orders are applied. Simulation results reveal that our proposed single branch GL-SIC significantly outperforms the linear MMSE receiver, the conventional SIC and exceeds the performance of MB-SIC with $L_b=4$ and MB MF-SIC with $L_b=4$ for the same BER performance. \vspace{-0.25cm} \begin{figure}[!htb] \centerline{ \includegraphics[width=0.18\textwidth]{Fig4a.eps} \includegraphics[width=0.18\textwidth]{Fig4b.eps} }\vspace{-0.55em}\caption{\footnotesize a)\ BER versus SNR for uplink cooperative system (N=16)\ \ \ b)\ BER versus number of users for uplink cooperative system (N=16)} \label{fig4} \vspace{-1em} \end{figure} The second example denotes the cooperative system where relays assist to process and forward the information. Fig. \ref{fig4}(a) shows the BER versus SNR plot for the different multi-relay selection strategies, where we apply the GL-SIC $(L_b=1)$ detection scheme at both the relays and the destination in an uplink cooperative scenario with 10 users and 6 relays. The performance bound for exhaustive search is presented here for comparison, where it examines all possible relay combinations and picks the best one with the highest SINR. From the results, it can be seen that with relay selection, the BER performance substantially improves. Furthermore, the BER performance curve of our proposed algorithm outperforms the standard greedy algorithm and approaches to the same level of the exhaustive search, whilst keeping the complexity reasonably low for practical utilization. The algorithms are then assessed in terms of the BER versus number of users in Fig. \ref{fig4}(b) with a fixed SNR=15dB. Similarly, we apply the GL-SIC $(L_b=1)$ detector at both the relays and the destination. The results indicate that the overall system performance degrades as the number of users increases. It also suggests that our proposed greedy relay selection method is more suitable than the standard greedy relay selection and non-relay selection scenario with a high SNR and a large number of users, as its curve does not change considerably when the number of users increases. \vspace{-0.3cm} \begin{figure}[!htb] \begin{center} \def\epsfsize#1#2{0.8\columnwidth} \epsfbox{Fig5} \vspace{-1.55em}\caption{\footnotesiz BER versus SNR for uplink cooperative system with different filters employed in the relays and the destination (N=16)} \label{fig5} \end{center} \vspace{-1.5em} \end{figure} In order to further verify the performance for the proposed cross-layer design, we compare the effect of different detectors with 10 users and 6 relays when this new greedy multi-relay selection algorithm is applied in the system. The results depicted in Fig. \ref{fig5} indicate that the GL-SIC approach allows a more effective reduction of BER and achieves the best performance that is quite close to the single user scenario, followed by the MB MF-SIC detector, the MB-SIC detector, the linear MMSE receiver and the conventional SIC with RAKE receivers employed at each stage. \vspace{-0.5em} \vspace{-0.35cm} \section{conclusion}\vspace{-0.2em} \vspace{-0.35cm} A cross-layer design strategy that incorporates the greedy list-based joint successive interference cancellation (GL-SIC) detection technique and a greedy multi-relay selection algorithm for the uplink of cooperative DS-CDMA systems has been presented in this paper. This approach effectively reduces the error propagation generated at the relays, avoiding the poorest relay-destination link while requiring a low complexity. Simulation results demonstrated that the proposed cross-layer optimization technique is superior to the existing techniques and can approach the exhaustive search very closely. \vspace{-0.35cm} {\footnotesize \bibliographystyle{IEEEbib}
\subsubsection{\@startsection{subsubsection}{3}{10pt}{-1.25ex plus -1ex minus -.1ex}{0ex plus 0ex}{\normalsize\bf}} \def\paragraph{\@startsection{paragraph}{4}{10pt}{-1.25ex plus -1ex minus -.1ex}{0ex plus 0ex}{\normalsize\textit}} \renewcommand\@biblabel[1]{#1} \renewcommand\@makefntext[1 {\noindent\makebox[0pt][r]{\@thefnmark\,}#1} \makeatother \renewcommand{\figurename}{\small{Fig.}~} \sectionfont{\large} \subsectionfont{\normalsize} \fancyfoot{} \fancyfoot[LO,RE]{\vspace{-7pt}\includegraphics[height=9pt]{LF}} \fancyfoot[CO]{\vspace{-7.2pt}\hspace{12.2cm}\includegraphics{RF}} \fancyfoot[CE]{\vspace{-7.5pt}\hspace{-13.5cm}\includegraphics{RF}} \fancyfoot[RO]{\footnotesize{\sffamily{1--\pageref{LastPage} ~\textbar \hspace{2pt}\thepage}}} \fancyfoot[LE]{\footnotesize{\sffamily{\thepage~\textbar\hspace{3.45cm} 1--\pageref{LastPage}}}} \fancyhead{} \renewcommand{\headrulewidth}{1pt} \renewcommand{\footrulewidth}{1pt} \setlength{\arrayrulewidth}{1pt} \setlength{\columnsep}{6.5mm} \setlength\bibsep{1pt} \twocolumn[ \begin{@twocolumnfalse} \noindent\LARGE{\textbf{Evaporative Deposition in Receding Drops$^\dag$}} \vspace{0.6cm} \noindent\large{\textbf{Julian Freed-Brown}}\vspace{0.5cm} \noindent\textit{\small{\textbf{Received Xth XXXXXXXXXX 20XX, Accepted Xth XXXXXXXXX 20XX\newline First published on the web Xth XXXXXXXXXX 200X}}} \noindent \textbf{\small{DOI: 10.1039/b000000x}} \vspace{0.6cm} \noindent \normalsize{We present a framework for calculating the surface density profile of a stain deposited by a drop with a receding contact line. Unlike a pinned drop, a receding drop pushes fluid towards its interior, continuously deposits mass across its substrate as it evaporates, and does not produce the usual ``coffee ring.'' For a thin, circular drop with a constant evaporation rate, we find the surface density of the stain goes as $\eta(r) \propto \left(\left(r/a_0\right)^{-1/2}-r/a_0\right)$, where $r$ is the radius from the drop center and $a_0$ is the initial outer radius. Under these conditions, the deposited stain has a mountain-like morphology. Our framework can easily be extended to investigate new stain morphologies left by drying drops.} \vspace{0.5cm} \end{@twocolumnfalse} ] \footnotetext{\dag~Electronic Supplementary Information (ESI) available: [details of any supplementary information available should be included here]. See DOI: 10.1039/b000000x/} \footnotetext{\textit{Department of Physics and the James Franck Institute, University of Chicago, 929 E 57th Street, Chicago, IL 60637, USA. \\E-mail: <EMAIL>}} \section{Introduction} Solute deposition from evaporating sessile drops is an important tool with varied applications. Evaporation-controlled deposition is used in colloidal self assembly,\cite{abkarian2004colloidal, nobile2009self, byun2010hierarchically, marin2011order, li2013macroscopic} electronics, \cite{kim2006direct, park2014flexible} particle segregation, \cite{monteux2011packing} and medical physics.\cite{brutin2011pattern} Currently studied effects on evaporative deposition include Marangoni flow,\cite{hu2006marangoni} substrate shape,\cite{xu2007evaporation, hong2009robust} and surface-bound colloids.\cite{yunker2011suppression} Here, we examine the impact of a receding contact line on solute deposition in an evaporating sessile drop. Evaporation and changes in height force fluid flow within a drying drop and eventually govern the shape of its deposited stain. Often, the outer edge of the drop pins to the surface, giving rise to the ``coffee ring effect,'' \cite{deegan1997capillary, deegan2000contact} but there are many cases when the edge does not stay pinned throughout evaporation. Many studies examine the connection between stain morphology and the stick-slip behavior of the contact line. \cite{deegan2000pattern, yabu2005preparation, maheshwari2008coupling, frastia2011dynamical, mampallil2012control, zhang2014coffee} Furthermore, a recent study by Li \emph{et al}\cite{li2014solute} found that drops containing poly(ethylene glycol) recede during the majority of their evaporation and only show contact line pinning at early times. These drops form mountain-like deposits in the center of the drop instead of the usual ring morphology. Other experiments have also found unusual, mountain-like stain morphologies for receding drying drops.\cite{willmer2010growth, baldwin2012monolith} In these cases, the standard intuition from the coffee ring effect does not apply. When the drop's edge freely recedes, height changes near the edge of the drop are larger than height changes at the center, which pushes fluid radially inward as the drop evaporates. Furthermore, mass deposits when it reaches the receding edge of the drop, which means mass continuously deposits across the surface instead of the original outer edge. Despite these differences, many of the ideas from the coffee ring effect can be extended to drops with receding contact lines. In this paper, we theoretically examine solute deposition in a thin, circular drop with a constant evaporation profile across its surface and a receding contact line. For this specific case, the surface density of the stain can be calculated analytically. To find the surface density of the deposited stain we \begin{enumerate} \item find the height profile as a function of time; \item determine the fluid velocity from evaporation and changes in height; \item solve for the trajectory of fluid parcels; and \item track the evolution of masses bound by fluid parcels. \end{enumerate} We find an exact formula for the surface density of the stain deposited from uniform evaporation. Our calculation serves as an illustrative example of the effects from a receding contact line and our method can easily be implemented numerically to handle more complex evaporation profiles. \section{Theoretical Regime} In our calculation, we consider small, circularly symmetric drops with slow dynamics. In these drops, surface tension dominates over gravitational, viscous, and inertial stresses. The drop height evolves quasi-statically and surface tension alone governs its shape. For drops with viscosity, density, and surface tension comparable to water, this regime corresponds to drop radii of a few millimeters and drying times of thousands of seconds. Note that these scales compare with scales found in experiments.\cite{deegan2000contact, li2014solute} The drop surface meets the substrate with a contact angle $\theta_c$. Moving contact lines are generally characterized by an advancing and a receding contact angle controlled by substrate heterogeneity \cite{drelich1996effect} and evaporating drops are often observed to recede at fixed contact angles.\cite{li2014solute} Based on these observations, we examine the case where $\theta_c$ is constant. We also work in the thin drop limit, where $\theta_c$ is small. In the thin drop limit, the mean curvature of the drop height is constant. Additionally, vertical velocities are negligible compared to radial velocities. For very thin drops, vertical length scales are much smaller than horizontal length scales and the diffusion length of solute over short times can be larger than typical vertical distances but smaller than typical horizontal distances. Then, we can assume that solute stays completely mixed in the vertical direction while horizontal diffusion is negligible. Furthermore, we adopt the limit where the solute is a dilute (its initial concentration, $\phi_0$, is small) and non-interacting. Under these assumptions, the solute acts as a passive tracer and is simply transported with the fluid's depth averaged velocity. We can neglect stratifying effects and any jamming or shock fronts that could be caused by solute interaction, which greatly simplifies the analysis. These assumptions were used by Deegan \emph{et al}\cite{deegan2000contact} to make useful predictions for pinned drops. We adopt the same conditions to facilitate comparison between the pinned and unpinned cases. \section{Drop Evolution} The fluid velocity within the drop determines the final surface density of the deposited stain. The depth averaged velocity obeys a continuity equation because the mass of fluid is conserved. \begin{figure}[t] \centering \includegraphics[width=.4\textwidth]{continuity.pdf} \caption{ Radial vertical section of droplet indicated by black line representing height, $h$, vs radial distance, $r$. Surface evaporation $(J)$ and changes in height $(\partial h/\partial t)$ induce flow ($u$) within drop. } \label{fig:continuity} \end{figure} Consider an annulus of fluid within the drop (See Fig.~\ref{fig:continuity}). As time evolves, the height profile decreases and acts as a local source of fluid flow $(\partial h/\partial t)$. Meanwhile, fluid leaves the surface through evaporation, which acts as a local sink. The evaporation rate, $J(r,t)$, is the volume of fluid removed per unit surface area per unit time at radius $r$ and time $t$. Together, evaporation and the change in height act as a source in a continuity equation for the velocity field: \begin{equation} \nabla\cdot(h u) = -\frac{\partial h}{\partial t}- \left(1+\left(\frac{\partial h}{\partial r}\right)^2\right)^{1/2} J(r,t), \end{equation} where $hu$ is the fluid flux, $\partial{h}/\partial{t}$ is the change in height over time, and $(1+(\partial h/\partial r)^2)^{1/2}$ is the surface area element. In the thin drop limit, $\frac{\partial h}{\partial r} \ll1$ and this equation simplifies to \begin{equation} \nabla\cdot(h u)=-\frac{\partial h}{\partial t} - J(r,t). \label{eq:continuity} \end{equation} To find the fluid velocity, we first use surface tension to find the height profile at a given time. By balancing the total volume change of the height profile against the total evaporation rate, we also solve for the time dependence of the height profile. From there, we substitute $h$ into Eq.~2 to solve for the velocity. \subsection{Height Profile} Surface tension determines the shape of the drop's surface. In the thin drop limit, the mean curvature of the surface $(\nabla^2 h)$ is constant and the drop is assumed to have a constant contact angle with the surface, $\theta_c$, as it evaporates. Together, these constraints determine the shape of the drop, but not its time dependence. In the thin drop limit, the height of an axisymmetric drop is given by \begin{equation} h(r,t) = H(t) \left(1-\frac{r^2}{a^2(t)}\right), \label{eq:h} \end{equation} where the height at the center of the drop, $H(t)$, and the maximum radius of the drop, $a(t)$, are both functions of time. Furthermore, $H(t)$ is proportional to $a(t)$ because the drop's contact angle with the surface is constant---that is, $\left.\frac{\partial h}{\partial r}\right|_{r=a(t)}$ is constant. While $H(t)$ and $a(t)$ have some unknown time dependence, they vary in time together (See Fig.~\ref{fig:height}). \begin{figure}[t] \centering \includegraphics[width=.4\textwidth]{hprofile.pdf} \caption{ Height profile at two times, showing time evolution of height profile in evaporating drop. If height profile shifts vertically (dotted line), contact angle decreases. Fluid must retract as height decreases to maintain a constant contact angle. } \label{fig:height} \end{figure} Evaporation from the drop's surface determines the exact form of the time dependence. The total evaporation must exactly balance the drop's volume change since evaporation conserves the mass of the fluid. Therefore, \begin{equation} \frac{dV}{dt} = -\int_0^{a(t)} J(r,t)2\pi r\ dr, \label{eq:balance} \end{equation} where the volume, $V$, is determined by the height profile: $V=\frac{1}{2} \pi H(t) a^2(t)$. Note that Eq.~4 holds for any $J(r,t)$. In the case where the evaporation rate is constant, Eq.~4 implies \begin{equation} \pi a(t) H(t) \dot a(t)+\frac{1}{2}\pi a^2(t) \dot H(t) = -\pi a^2(t) J_0, \end{equation} where $J(r,t)=J_0$. This differential equation is solved by employing the fact that $H(t)\propto a(t)$. Then, \begin{equation} a(t) = a_0 \left(1-\frac{t}{t_f}\right) \label{eq:a} \end{equation} and \begin{equation} H(t) = H_0 \left(1-\frac{t}{t_f}\right), \end{equation} where $a_0$ is the initial radius of the drop, $H_0$ is the initial height at the center, and $t_f = \frac{3 H_0}{2 J_0}$ is the final drying time. \subsection{Fluid Velocity} Together, the height profile and continuity equation uniquely determine the depth averaged velocity profile, $u(r,t)$. After substituting $h(r,t)$ into Eq.~2, the velocity is given by \begin{equation} u(r,t) =\left\{ \begin{array}{lr} - \frac{r}{4(t_f-t)} & r\leq a(t)\\ 0 & r>a(t) \end{array} \right. \end{equation} The functional form of $u(r,t)$ immediately provides interesting results. In contrast to the pinned drop case, $u$ is inward and has a maximum value which is independent of time. At any time, the maximum velocity of a fluid parcel in the drop is $u(a(t),t)=-a_0/(4t_f)$. This maximal fluid velocity is slower than the rate that the edge recedes, $\dot a(t) = -a_0/t_f$. Therefore, the edge of the drop overtakes every fluid parcel that originates at a non-zero radius before the drop finishes drying. After a fluid parcel is overtaken, its mass deposits onto the substrate and remains immobile throughout the remainder of the evaporation. This deposition mechanism is qualitatively different from that for pinned drops. \section{Surface Density Profile} \begin{table}[tb] \begin{tabular}{c p{.4\textwidth}} \hline\\ Symbol & Definition\\ \hline \\ $R$ & Final deposition radius and Lagrangian label for a mass of solute\\ $M(R)$ & Mass bound by inner radius $R$ after deposition is complete\\ $\tau(R)$ & Deposition time; $a(\tau(R))=R$\\ $\mu(R)$ & Initial radius bounding $M(R)$ \\ $\xi(R,t)$ & Trajectory of bounding radius over time; $\xi(R,0)=\mu(R)$ and $\xi(R,\tau(R))=R$\\ $u(r,t)$ & Radial velocity of fluid at radius $r$ and time $t$\\ $\eta(r)$ & Deposited surface density profile\\ \hline \end{tabular} \caption{Definitions of symbols} \label{tab:def} \end{table} The evolution of the initial mass of solute determines the final surface density profile, $\eta(r)$. After the drop dries, the mass $M(R)$ deposited on the substrate between a radius $R$ and the initial edge of the drop $a_0$ is the integral over the surface density: \begin{equation} M(R) = \int_R^{a_0} 2\pi r\eta(r)dr, \end{equation} so that $M'(R)\equiv 2\pi R \eta(R)$. At time $t=0$, $M(R)$ can also be calculated from the initial height profile. If $\phi_0$ is the initial density of solute and $\mu(R)$ is the radius that initially bounds the mass $M(R)$, \begin{equation} M(R) = \int_{\mu(R)}^{a_0} 2\pi r \phi_0h(r,0)dr. \end{equation} (See Fig.~\ref{fig:deposition}) Then, $M'(R)$ explicitly connects $\eta$, $h$, and $\mu$: \begin{equation} \eta(R) =\phi_0h(\mu(R),0)\frac{\mu(R)}{R}\mu'(R). \label{eq:eta} \end{equation} Interpreted piece by piece, this is a very intuitive equation. The annular parcel of fluid that deposits at $R$ determines the final surface density. The parcel originated at $\mu(R)$ with a local area density of $\phi_0h(\mu(R),0)$. The circumference of the parcel decreases after being transported from $\mu(R)$ to $R$, leading to an increase in density by a factor of $\mu(R)/R$. The annular parcel is also be compressed or extended radially during its evolution, which further alters its density by a factor of $\mu'(R)$. Therefore, $\mu(R)$ determines the final surface density. \begin{figure}[htbp] \centering \begin{tabular}{|c|} \hline \includegraphics[width=.375\textwidth]{recedinga}\\ \hline \includegraphics[width=.375\textwidth]{recedingb}\\ \hline \includegraphics[width=.375\textwidth]{recedingc}\\ \hline \includegraphics[width=.375\textwidth]{recedingd}\\ \hline \includegraphics[width=.375\textwidth]{recedinge}\\ \hline \end{tabular} \caption{ Schematic of solute deposition from evaporating drop with receding contact line. (a) through (e) shows deposition in reverse chronological order. (a) After drop has dried, all solute has deposited on the substrate. $M(R)$ is the total mass bound between radius $R$ and outer radius of the stain. (b) When the drop edge, $a(t)$, is less than $R$, the mass $M(R)$ has deposited on substrate and remains immobile throughout the remainder of the evaporation. (c) At time $\tau(R)$, $a(t)=R$ and the last parcel of $M(R)$ deposits onto substrate. (d) When $a(t) > R$, $M(R)$ continuously deposits across substrate as drop dries. Its bounding radius, $\xi(R,t)$ evolves over time. (e) Initial bounding radius, $\mu(R)=\xi(R,0)$, and height profile, $h(r,0)$, determine $M(R)$. } \label{fig:deposition} \end{figure} We track the evolution of an annular fluid parcel that deposits at $R$ backwards in time to find its initial radius. Let $\xi(R,t)$ be the trajectory of a fluid parcel that finally deposits at $R$. The parcel will deposit on the surface at some time $\tau(R)$. By definition, $\xi(R,\tau(R))=R$ at this time. Since a fluid parcel deposits after the receding edge overtakes it, the parcel deposits when $a(\tau(R))=R$. Substituting this condition into Eq.~6 yields \begin{equation} \tau(R) =t_f(1- R/a_0). \label{eq:tau} \end{equation} The solute moves with the local fluid velocity because it is passively transported by the fluid. Thus, the trajectory of a fluid parcel is \begin{equation} \dot \xi(R,t) = u(\xi(R,t), t) = -\frac{\xi(R,t)}{4(t_f-t)}, \label{eq:dxi} \end{equation} so that $d\xi/\xi = -dt/(4(t_f-t)),$ implying $\xi \propto (1-t/t_f)^{1/4}$. Because the parcel reaches the edge at radius $R$ and time $\tau(R)$, its trajectory is subject to the boundary condition $\xi(R,\tau(R)) = R.$ Then, \begin{equation} \xi(R,t) = a_0 \left(\frac{R}{a_0}\right)^{3/4} (1-t/t_f)^{1/4} \end{equation} and \begin{equation} \mu(R)=\xi(R,0)=a_0 \left(\frac{R}{a_0}\right)^{3/4} \label{eq:mu} \end{equation} As shown in Eq.~11, the surface density is completely determined by $\mu$. Therefore, \begin{equation} \eta(r) = \frac{3}{4} \phi_0 H_0 \left(\left(\frac{r}{a_0}\right)^{-1/2}-\frac{r}{a_0}\right).\label{eq:analyticresult} \end{equation} See Fig.~\ref{fig:eta} for a plot of $\eta(r)$. The nature of the divergence at small $r$ can be found quite generally. As an example, we compute the asymptotic behavior of $\eta$ for any evaporation profile $J(r,t)$ that is only a function of $r/a(t)$. In these cases $u(r,t)$ varies linearly near the center of the drop as $u(r,t) \approx \nu r/(t_f-t)$, where the dimensionless coefficient $\nu$ depends on $J$ and $h$ as discussed below. Then, Eq.~13 implies that $\xi(R,t) \approx a_0 (R/a_0)^{1-\nu} (1-t/t_f)^{-\nu}$, which further implies that $\mu\propto (R/a_0)^{1+\nu}$ and $\eta\propto (R/a_0)^{2\nu}$. \begin{figure}[tb] \centering \includegraphics[width=.4\textwidth]{etaprofile2} \includegraphics[width=.4\textwidth]{surfacedensity3D2} \caption{ (a) Deposited surface density profile (blue, solid) compared with $\phi_0 h(r,0)$, the initial surface density (yellow, dashed). Note that the mass is concentrated and the surface density diverges at the center. The stain is shaped like a mountain instead of a ring. (b) Perspective view of the deposited stain. } \label{fig:eta} \end{figure} \section{Discussion} The first thing to note about the surface density profile (Eq.~16) is that mass is concentrated at the center and fades towards the outside of the drop. The overall shape is more like a mountain than a ring. In fact, the surface density diverges as $(r/a_0)^{-1/2}$ towards the center. This is a weak divergence and the total mass of the stain, $M(0)$, is finite. For experiments, the divergence is cut off because one of the assumptions made in this calculation will fail at late times. For example, at late times the concentration of the solute in the fluid will increase until the solute starts interacting with itself. After this point, the assumption that the solute is a dilute passive tracer is no longer valid and the divergence will not be realized. However, the exact time when the calculation fails depends on the experimental realization. For the given example, lowering the initial solute concentration causes the assumptions to remain valid for longer times. Remarkably, the power-law governing the deposit shape is controlled by experimental conditions. As noted above, only the behavior of $u(r)$ near the center is relevant for determining the density profile $\eta(r)$. The exponent $\nu$ can be easily calculated from the evaporation and height profiles without explicitly solving for the complete flow field: \begin{equation} \nu=\frac{1}{2}\left(1-\frac{J(0)/\bar{J}}{\dot{h}(0)/\dot{\bar{h}}}\right), \end{equation} where the overbar indicates the average over the drop.\cite{witten2009robust} This formula depends only on the existence of a stationary and regular stagnation point; thus it is applicable to unpinned as well as pinned circular drops. The unpinned aspect only influences the $\dot h(0)/\dot{\bar{h}}$ factor (See Fig.~\ref{fig:height}). When $\nu>0$, fluid flows away from the center and the surface density will fade to 0. For $\nu<0$, the density diverges and the stain forms a mountain. For a drop evaporating on a dry substrate, $\nu$ can be calculated analytically. The evaporation profile for a drop on a thin dry substrate is $J(r,t)=J_0 f(\lambda) \left(1-\left(r/a(t)\right)^2\right)^{-\lambda}$, where $\lambda = (\pi-2\theta_c)/(2\pi-2\theta_c)$.\cite{deegan2000contact} In the limit where $\theta_c$ approaches 0, $\nu =1/8$. In this case, the stain fades at the center even though the contact line recedes. It is also worth noting that Eqs.~2, 3, 4, and 11 are generic for a thin drop with a receding contact line. Even though the functional form of the evaporation controls the velocity profile, the time dependence of the height profile, and the final form of the surface density, it is possible to follow the procedure outlined in this paper to find the surface density. In special cases, the surface density can be found analytically, but it is not difficult to extend this analysis numerically to explore the morphology induced by other evaporation profiles. \section{Conclusion} The fluid motion within a drying drop is directly influenced by the behavior of its contact line. Unlike a pinned drop, a receding contact line pushes fluid inwards and the stain deposits continuously as the drop evaporates. For a circular drop with Poisson ratio $\nu$ at its center, the power law of the surface density profile near the center is $\eta\propto (r/a_0)^{2\nu}$. To apply this framework accurately to an experiment requires an accurate estimate of the surface evaporation. It is not clear, for example, whether deposited mass will retain moisture, which would greatly reduce the evaporation from the edges of the drying drop. Once an evaporation profile is obtained, our method can quickly predict the profile of the deposited stain. \section*{Acknowledgments} The author is grateful to Efi Efrati for fruitful discussions. This work is a PhD research project supervised by T. A. Witten. It was supported in part by the University of Chicago MRSEC program of the US National Science Foundation under Award Number DMR 0820054. \footnotesize{
\section{Introduction}\label{sec:Introduction} In this work we study the long time behavior of Markov processes, primarily in continuous time, whose states assign either $+1$ or $-1$, called a spin value, to each vertex of the $d$-dimensional lattice $\mathbb{Z}^d$. We will discuss two types of processes, a much studied one denoted $\sigma(t)= (\sigma_x (t):x\in \mathbb{Z}^d)$ and then a modified one, denoted $\sigma'(t)$, in which some vertices are ``frozen'' -- that is, their spin values are not allowed to change. Before giving complete definitions of $\sigma(t)$ and $\sigma'(t)$, we give brief descriptions and motivations. $\sigma(t)$ has an energy-lowering dynamics with energy of the form $ -\sum^* \sigma_x \sigma_y$, where $\sum^*$ denotes the sum over nearest neighbor pairs of vertices. Here, energy is lowered at the update of $\sigma_x$ if its value is changed to agree with a strict majority of neighbors. The modified process $\sigma'(t)$ basically corresponds to $\sigma(t)$ in a random environment where randomly or deterministically selected vertices are frozen from time zero, some plus and some minus. There are two distinct motivations for studying $\sigma'$. One, explained in more detail below, comes from the usual $\sigma$ process, with random initial state, but on a slab, say $\mathbb{Z}^2 \times \{0, \ldots, k-1 \}$, rather than on $\mathbb{Z}^d$. Here, random rectangular regions of the form $R \times \{0, \ldots, k-1 \}$ are fixed (thus effectively frozen) from time zero if they start with constant spin value. A second motivation comes from the energy-lowering dynamics of random-field models with energy $-\sum^* \sigma_x \sigma_y -\sum_x h_x \sigma_x$ where the $h_x$'s are i.i.d. variables with common distribution $\rho$. Suppose, for example, that $\rho = \rho^+ \delta_H + \rho^- \delta_{-H} + (1-\rho^+ -\rho^-)\delta_0$ with $H>2d$, where $\delta_r$ denotes the unit point measure at $r$. The reader can then check that at vertex $x$ with $h_x=H$ (resp., $-H$), either from time zero or else after the first update at $x$, $\sigma_x (t)$ will be fixed plus (resp., minus) and thus effectively frozen. We proceed now to define our two processes, $\sigma$ and $\sigma'$, and review some known results about~$\sigma$. \section{The two processes} \subsection{The process $\sigma$} The stochastic process $\sigma(t) = \sigma (t, \omega)$, where $\sigma_x (t)$ denotes the value of the spin at vertex $x \in \mathbb{Z}^d$ at time $t \geq 0$, starts from a random initial configuration $\sigma (0) =\{\sigma_x (0)\}_{x \in \mathbb{Z}^d}$, drawn from an independent Bernoulli product measure \begin{equation}\label{mu} \mu_{\theta} (\sigma_x (0) = + 1) = \theta = 1 - \mu_{\theta} (\sigma_x (0) = - 1). \end{equation} \noindent The system evolves in continuous time according to an agreement inducing dynamics: at rate 1, each vertex changes its value if it disagrees with more than half of its neighbors, and decides its spin value by tossing a fair coin in the event of a tie. This process corresponds to the zero-temperature limit of Glauber dynamics for a stochastic Ising model with ferromagnetic nearest-neighbor interactions and no external magnetic field (see, for example, \cite{NNS} or \cite{KRB}), having Hamiltonian (energy function) \begin{equation}\label{hamiltonian} \mathcal{H} = - \sum_{\{x, y\}: \|x - y\| = 1} \sigma_x \sigma_y , \end{equation} \noindent where $\|x\|$ denotes the Euclidean norm of $x$. More precisely, let $\mathcal{S}$ be the state space of configurations $\sigma$, i.e., $\mathcal{S} =\{- 1, 1\}^{\mathbb{Z}^d}$. The continuous time dynamics can be defined by means of independent, rate 1 Poisson processes (clocks), one assigned to each vertex $x$. If the clock at vertex $x$ rings at time $t$ and the change in energy \begin{equation*} \Delta \mathcal{H}_x (\sigma) = 2 \sum_{y : \|x - y\|= 1} \sigma_x \sigma_y \end{equation*} \noindent is negative (respectively, positive), a spin flip (that is, a change of $\sigma_x$) is done with probability 1 (respectively 0). To resolve the case of ties when $\Delta \mathcal{H}_x (\sigma) = 0$, each clock ring is associated to a fair coin toss and a spin flip is done with probability 1/2, or equivalently $\sigma_x$ is made to be $+1$ (respectively, $-1$) if the coin toss comes up heads (resp., tails). Let $\mathbb{P}_{\mathrm{dyn}}$ be the probability measure for the realization of the dynamics (clock rings and tie-breaking coin tosses), and denote by $\mathbb{P}_\theta= \mu_{\theta} \times \mathbb{P}_{\mathrm{dyn}}$ the joint probability measure on the space $\Omega$ of initial configurations $\sigma (0)$ and realizations of the dynamics; an element of $\Omega$ is denoted~$\omega$. This process has been studied extensively in the physics and mathematics literature -- primarily on graphs such as the hyperlattice $\mathbb{Z}^d$ and the homogeneous tree of degree $K$, $\mathbb{T}_K$. A physical motivation, which corresponds to the symmetric initial spin configuration, is the behavior of a magnetic system following a deep quench. A deep quench occurs when a system that has reached equilibrium at an initial high temperature $T_1$ is instantaneously subjected to a very low temperature $T_2$. Here we take $T_1 =\infty$ and $T_2 =0$. For references on this and related problems see, for example, \cite{NNS} or \cite{KRB}. The main focus in the study of this model is the formation and evolution of boundaries delimiting same spin cluster domains. These domains shrink or grow or split or coalesce as their boundaries evolve. This model is often referred to as a model of \emph{domain coarsening}. A fundamental question is whether the system has a limiting configuration, or equivalently does every vertex eventually stop flipping? Whether \begin{equation}\label{sigmainfinity} \lim_{t\rightarrow\infty}\sigma_x(t) \end{equation} \noindent exists for almost every initial configuration, realization of the dynamics and for all $x\in \mathbb{Z}^d$ depends on $\theta$ and on the dimension $d$. We refer to the existence of the limit~\eqref{sigmainfinity} as {\bf fixation} at $x$. Nanda, Newman and Stein \cite{NNS} investigated this question when $d=2$ and $\theta=\frac{1}{2}$ and found that the limit does not exist; that is, every vertex flips infinitely often. Their work extended an old result of Arratia \cite{A}, who showed the same happens on $\mathbb{Z}$ (for $\theta\neq 0$ or $1$). It is an open problem to determine what happens for $d\geq 3$ and $\theta = 1/2$. One important consequence of the methods of \cite{NNS} is that if each vertex of the graph has odd degree (for example, on $\mathbb{T}_K$ for $K$ odd), then $\sigma_x (\infty)$ does exist for almost every initial configuration, realization of the dynamics and every vertex $x$. Another question of interest is whether sufficient bias in the initial configuration leads the system to reach consensus in the limit. That is, does there exist $\theta_{\ast}\in(0, 1)$, such that for $\theta \geq \theta_{\ast}$, \begin{equation}\label{fixation} \forall x \in G, \mathbb{P}_{\theta} (\exists T = T (\sigma (0), \omega, x) < \infty \text{ so that } \sigma_x (t) = + 1 \text{ for } t \geq T) = 1? \end{equation} \noindent We will refer to \eqref{fixation} as {\bf fixation to consensus}. It was conjectured by Liggett \cite{L} that fixation to consensus holds for all $\theta > \frac{1}{2}$. Fontes, Schonmann and Sidoravicius \cite{FSS} proved fixation to consensus for all $d\geq 2$ and $\theta_{\ast}$ strictly less but very close to 1. On $\mathbb{T}_3$, however, Howard \cite{Howard} showed that for some $\theta>1/2$, the system does not fixate to $+1$ consensus. \subsection{$\sigma$ on slabs} In \cite{DKNS1} and \cite{DKNS2}, Damron, Kogan, Newman and Sidoravicius studied coarsening started from a configuration sampled from $\mu_\theta$ on two-dimensional slabs of finite width $k$ with free boundary conditions, which we denote as $\text{Slab}_k$. These are graphs with vertex set $\mathbb{Z}^2 \times \{0, 1, \ldots, k-1\}$ ($k \geq 2$) and edge set $\mathcal{E}_k =\{\{x, y\}: \|x-y\|_1=1\}$. Their work was motivated by the question of whether there are vertices that fixate for $d\geq 3$ (and for which values of $d$). It has been suggested by computational results of Spirin, Krapivsky and Redner \cite{SKR} that some vertices do indeed fixate. The work in \cite{DKNS1} and \cite{DKNS2} on slabs highlights the possible diferences in long time behavior between $\mathbb{Z}^2$ and $\mathbb{Z}^3$. The authors showed that if $k=2$ the system fixates with both free and periodic boundary conditions; if $k=3$ with periodic boundary conditions the system also fixates; for all $k\geq 4$ with periodic boundary conditions some vertices fixate for large times and some do not, and the same result holds for all $k\geq 3$ with free boundary conditions. We call vertices which change spin sign forever flippers. One question which remains open, is whether the set of flippers percolates (contains an infinite component). On $\text{Slab}_3$, each $v\in \mathbb{Z}^2 \times\{0\}$ or $v \in \mathbb{Z}^2 \times\{2\}$ has five neighbors, so by a variant of the arguments of Nanda, Newman, Stein \cite{NNS}, $v$ fixates almost surely. Therefore on this graph, flippers can only exist in $\mathbb{Z}^2 \times \{1\}$. On the other hand, if the initial configuration on $\text{Slab}_3$ is chosen according to a symmetric Bernoulli product measure, then by the Ergodic Theorem there are at $t=0$ infinitely many pillar-like same-spin formations (say, $2\times 2 \times 3$ blocks) that are stable under the dynamics. These pillars are analogous to frozen vertices on $\mathbb{Z}^2$ in our new process $\sigma'$. A general version of this new process on $\mathbb{Z}^d$ is presented in the following section. \subsection{The process $\sigma'$} In this subsection we define a new stochastic process on $\mathbb{Z}^d$, which we denote by~$\sigma'(t)$, in which some vertices are frozen for all time and the others are not. There are two basic versions of $\sigma'$, which we will call {\bf disordered} and {\bf engineered}, according to whether the frozen vertices are chosen randomly or deterministically. The initial configuration of the disordered $\sigma'$ will be assigned as follows. Fix $\rho^+, \rho^- \geq 0$ with $\rho^+ + \rho^- \leq 1$ and pick three types of vertices (frozen plus, frozen minus and unfrozen) by i.i.d. choices with respective probabilities $\rho^+, \rho^-$ and $1-(\rho^+ + \rho^-)$. Once the frozen vertices have been chosen and assigned a spin value, the non-frozen vertices will be assigned spin values arbitrarily. (In other words, the theorems we prove will be valid for all choices of such spin values.) We will denote by $\mathbb{P}=\mathbb{P}_{\rho^+, \rho^-} \times \mathbb{P}_{\text{dyn}}$ the overall probability measure where $\mathbb{P}_{\rho^+, \rho^-}$ is the distribution for the assignment of frozen plus and minus vertices and $\mathbb{P}_{\text{dyn}}$ is the distribution of the following dynamics for $\sigma'(t)$. The continuous time dynamics is defined similarly to that of the $\sigma(t)$ process. Vertices are assigned independent, rate 1 Poisson clock processes and tie-breaking fair coins, and flip sign to agree with a majority of their neighbors. Frozen vertices, however, never flip regardless of the configuration of their neighbors. In the engineered $\sigma'$ the frozen vertices are chosen deterministically while the non-frozen vertices are assigned values at time zero by i.i.d. choices with respective probabilities $\theta$ and $1-\theta$ for $+1$ and $-1.$ As usual, we are interested in the long term behavior of this model depending on the dimension $d$, the densities of frozen vertices, $\rho^+, \rho^-$ (or in the engineered version, the choice of frozen vertices), and the initial configuration of vertices, which we denote by $\sigma'(0)$. $\sigma'(0)$ may be regarded as an element of $\{-1,+1\}^{\mathbb{Z}^d}$ even though the frozen vertex spins are instantaneously replaced (at time $t=0+$) by their frozen values. Note that the $\sigma'$ processes for all possible choices of $\sigma'(0)$ are coupled on a single probability space. When $\rho^+>0 $ and $\rho^->0$, almost surely there exist flippers. To see this, consider the following configuration for the case $d=2$, which has probability $(1-\rho^+-\rho^-)(\rho^+)^2(\rho^-)^2$: the vertex labeled $v$ in Figure~\ref{fig:flippervertex} below is not frozen, but has two frozen neighbors of spin $+1$ and two frozen neighbors of spin $-1$, and thus flips infinitely often. Similar flippers, as well as more complicated clusters of flippers, occur for any $d$. \begin{figure}[h] \centering \includegraphics{Zdfrozen-1.pdf} \caption{A vertex that flips infinitely often.} \label{fig:flippervertex} \end{figure} We conclude that $\sigma_v (t)$ need not have a limit as $t \rightarrow \infty$. So we are interested in whether the flippers percolate. Of course, one may study percolation of any of the three types of vertices (fixed plus, fixed minus or flippers) or of the union of two of the three. Theorem \ref{thm_frozen_+-} below answers a question in this direction. \section{Main results}\label{sec:theorems} Our first two theorems concern the disordered version of $\sigma'$ and the last concerns engineered versions. The first theorem is a fixation to consensus result for the case of positive initial density of frozen $+1$'s and zero initial density of frozen $-1$'s. The second theorem is a more general result in which both $\rho^+, \rho^- >0$, but we require $\rho^-$ to be much smaller than~$\rho^+$. For this more general case we obtain that the set of flippers together with (eventually) fixed vertices of spin $-1$ does not percolate. \begin{theorem}\label{thm_frozen_+d} Consider the disordered stochastic process $\sigma'$ on $\mathbb{Z}^d$ for any $d$ and any $\rho^+ >0$, with $ \rho^- = 0$. Then $\mathbb{P}($the system fixates to~$+1$ consensus for any $\sigma'(0))=1$. \end{theorem} \begin{theorem}\label{thm_frozen_+-} Consider the disordered stochastic process $\sigma'$ on $\mathbb{Z}^d$ for any $d$ and any $\rho^+ >0$, with $\rho^- >0 $ sufficiently small (depending on $\rho^+$ and $d$). For an initial configuration $\sigma'(0)$, denote by $\mathcal{C}(\sigma'(0))$ the collection of (eventually) fixed $-1$ vertices and flippers. Then $\mathcal{C}$ does not percolate: \[ \mathbb{P}(\mathcal{C}(\sigma'(0)) \text{ contains an infinite component for some }\sigma'(0))=0. \] \end{theorem} \begin{remark}\label{remarkmodify} Our main results, with essentially the same proofs, remain valid when the process $\sigma'$ is modified in various ways. For example, the rules for breaking ties can be modified as long as there is strictly positive probability to update to $+1$. Also the process can evolve in discrete time with synchronous updating. Another modification is to replace the two spin values, $\{-1,+1\}$, by $q$ values, say $\{1,2,\dots,q\}$, as long as updates respect a majority agreement of neighbors on one value. \end{remark} The next theorem concerns engineered versions of $\sigma'$ in which the frozen vertices are all $+1$ and chosen deterministically while the other spin values at time zero are i.i.d. with probability $\theta$ to be $+1$. Since in the disordered $\sigma'$ there are infinitely many frozen vertices, it's natural to consider choosing infinitely many frozen $+1$ vertices forming a lower dimensional subset of $\mathbb{Z}^d$, such as $\mathbb{Z}^{d-k} \times \emptyset_k$ with $k, d-k \geq 1$, where $\emptyset_k$ denotes the origin in $\mathbb{Z}^k$. Although we have no results to report for any of these situations, the next theorem concerns a slab approximation to the $k=1$ case where all the vertices in a codimension one hyperplane are frozen to $+1$. Note that the frozen hyperplane separates $\mathbb{Z}^d$ into two graphs isomorphic to $\mathbb{Z}^{d-1} \times \{0, 1, 2, \ldots \}$ with $\mathbb{Z}^{d-1}\times \{0\}$ frozen to $+1$, that evolve independently of each other. \begin{theorem}\label{thm_slab} Consider the engineered stochastic process $\sigma'$ on $\mathbb{Z}^{d-1}\times \{0, 1, 2, \ldots, K \}$, for $d\geq~2$ and $K\geq 1$, with all spins on $\mathbb{Z}^{d-1} \times \{0\}$ frozen to $+1$ and $\sigma'(0)$ for other vertices chosen from $\mu_\theta$. Then for any $\theta>0$, with probability $1$ the system fixates to $+1$ consensus. \end{theorem} \section{Proofs}\label{sec:proofs} In this section we give the proofs of the theorems stated in Section~\ref{sec:theorems}. Let $B_L=[-L, L]^d$ be the cubic box of side length $2L+1$ centered at the origin and $B_L(x)=x + [-L, L]^d$ be the translated box centered at $x\in \mathbb{Z}^d$. \subsection{Bootstrap percolation} Following Fontes, Schonmann and Sidoravicius \cite[Section~2]{FSS}, we describe the bootstrap percolation process that assigns configurations $\{u, s\}^{\mathbb{Z}^d}$ to a subset of $\mathbb{Z}^d$; here $u$ represents an \emph{unstable} spin and $s$ represents a \emph{stable} spin at a vertex. \begin{definition} The $d$-dimensional $(u\rightarrow s)$ bootstrap percolation process with threshold $\gamma$, defined in a finite or infinite volume $\Lambda \subseteq \mathbb{Z}^d$, starting from the initial configuration $\eta(0) \in \{u, s\}^\Lambda$ is a cellular automaton which evolves in discrete time $t=0, 1, 2, \ldots$ such that at each time unit $t\geq 1$ the current configuration is updated according to the following rules. For each $x\in \Lambda$, \begin{enumerate} \item If $\eta_{x}(t-1)=s$, then $\eta_x(t)=s$. \item If $\eta_{x}(t-1)=u$, and at time $t-1$ the vertex $x$ has at least $\gamma$ neighbors in $\Lambda$ in state $s$, then $\eta_x(t)=s$; otherwise, the spin at vertex $x$ remains unchanged; that is, $\eta_x(t)=u$. \end{enumerate} \end{definition} We will consider this process with threshold $\gamma = d$, as its evolution is close to our coarsening dynamics, and assume the initial configuration to be chosen from an independent Bernoulli product measure $P(\eta_x(0)=s)=p$, for $p$ small, on $\Lambda = \mathbb{Z}^d$. Note that by monotonicity of the dynamics, each $\eta_x(t)$ has a limit as $t\to\infty$. \begin{definition} A configuration $\eta\in \{u, s\}^\Lambda$ \textbf{internally spans} a region $B_L(x) \subset \Lambda$, if the bootstrap percolation restricted to $B_L(x)$, started from $\eta |_{B_L}$, ends up with all vertices of $B_L$ in state $s$. We will denote by $\eta_L$ the subset $\{x \in B_L: \, \eta_x=s\}$ for such an $\eta$ and call it \textbf{spanning}. \end{definition} The following proposition, an immediate consequence of results of Schonmann \cite{S}, provides a key ingredient for our proofs. \begin{proposition}\label{schonmann} \emph{[Schonmann]} If $p>0$, then \begin{equation*} \lim_{L \rightarrow \infty} P(B_L \text{ is internally spanned}) = 1. \end{equation*} \end{proposition} \begin{remark} In \cite{S} a variation of bootstrap percolation with threshold $\gamma=d$, called th \\ {\bf {modified basic model}}, is considered. Here, rule 2 is replaced by one which requires at least one neighbor in each of the $d$ coordinate directions from $x$ to have value $s$ in order that $x$ change from $u$ to $s$. For the modified basic model, the analogue of Proposition~\ref{schonmann} remains valid -- see \cite[Proposition~3.2]{S}. This will be used in the proof of Theorem~\ref{thm_slab}. \end{remark} \subsection{Preliminary lemmas} We will make a comparison between the bootstrap percolation process on $\mathbb{Z}^d$ and our process $\sigma'$ by mapping frozen plus spins to stable spins $s$, and all other spins to unstable spins $u$. In fact, we will compare only on finite regions which do not contain frozen minus vertices, so it is unimportant that these vertices are mapped to $u$. We say that a region $B_L (x)$ contains a \emph{spanning subset} of frozen plus vertices if the configuration obtained by the above mapping internally spans $B_L(x)$. \begin{definition}\label{Mcaptured} $B_L(x)$ is \textbf{entrapped} if it contains a spanning subset of frozen plus vertices. It is \textbf{captured} if it is entrapped and all $2^d$ corners are frozen plus. It is \textbf{$M$-captured} ($M\in \{0, 1, \ldots, L\}$) if it is entrapped and for each of the $2^d$ corners $C^i, (i=1, \ldots, 2^d)$, and each coordinate direction $j=1, \ldots, d$ there is a frozen plus vertex within $B_L(x)$ of the form $C^{i, j}=C^i+m e^j$ with $|m| \leq M$ (where $e^1, \ldots, e^d$ are the standard basis vectors of $\mathbb{R}^d$) -- see Figure~\ref{fig:Mcaptured}. Note that captured is the same as 0-captured. \end{definition} \begin{figure}[h] \centering \includegraphics{Zdfrozen-2.pdf} \hspace{.75in} \caption{$B_L(x)$ is M-captured.} \label{fig:Mcaptured} \end{figure} The notion of $M$-captured will be used (in Lemma \ref{lemmaB[M]}) to guarantee that with high probability most of the vertices in the box $B_L$ (at least for $1 \ll M \ll L$) will fixate to +1. Note that, although in the above definition $M = L$ is allowed, in Lemmas \ref{lemmaprobMgood} and \ref{lemmaB[M]} we require $M <L$. Lemma \ref{lemmaMcaptured}, though, will allow us to choose $1 \ll M \ll L$ in the final proof, although we do not make use of this. Lemmas \ref{lemmaboostrapweak}, \ref{lemmabootstrap} and \ref{lemmacaptured}, presented next, will be used in the proof of Theorem \ref{thm_frozen_+d}. \begin{lemma}\label{lemmaboostrapweak} Given $L$ and a spanning subset $\eta_L$ of $B_L$, consider the $\sigma'$ process in $\mathbb{Z}^d$ with initial spins in $\eta_L$ taken as frozen plus and all others in $\mathbb{Z}^d \setminus \eta_L$ taken as minus but not frozen. Then \begin{equation*} \mathbb{P} (\text{for some } t\in[0, 1], \sigma'(t)|_{B_L} \equiv +1) >0. \end{equation*} \end{lemma} \begin{proof} Since $\eta_L$ is a spanning subset of $B_L$, the corresponding bootstrap percolation process occupies all vertices of $B_L$ in a finite number of steps. Since the threshold $\gamma=d$, we can, with a small but positive probability, arrange the clock rings for $t\in (0, 1)$ and tie-breaking coin tosses of the coarsening dynamics to mimic the dynamics of bootstrap percolation (in a much longer discrete time interval). Thus $\sigma'(t)|_{B_L} \equiv +1$ for some $t\in [0, 1]$ with positive probability. \end{proof} The next lemma strengthens the last one by showing that, if we allow the process to run until a large time, then with probability close to one all the vertices of $B_L$ will flip to $+1$ before that time. Let $Span_L$ be the event that the frozen plus vertices span $B_L$ and there are no frozen minus vertices in $B_L$. \begin{lemma}\label{lemmabootstrap} Given $L <\infty$, \begin{equation*} \lim_{T \to \infty} \mathbb{P} \left(\exists t\in[0, T] \text{ such that } \sigma'(t)|_{B_L} \equiv +1 \text{ for all }\sigma'(0) \mid Span_L \right) = 1. \end{equation*} \end{lemma} \begin{proof} Pick a maximal spanning subset $\eta_L$ (in some deterministic ordering of subsets) of frozen plus vertices of $B_L$. By Lemma \ref{lemmaboostrapweak} and attractiveness (monotonicity of the dynamics), there is an $\epsilon' >0$ such that, for any $m$ and $\sigma'(m)$ (consistent with the frozen vertices in $B_L$), \begin{equation*} \mathbb{P} (\text{for some } t\in[m, m+1], \sigma'(t)|_{B_L} \equiv +1 | \sigma'(m)) \geq \epsilon'. \end{equation*} \noindent Given $\epsilon>0$, let $T_L \geq \frac{\log \epsilon}{\log (1-\epsilon')}$ be an integer and apply repeatedly the last inequality to time intervals of the form $[m, m+1]$ for integers $m \in [0,T_L)$: \[ \mathbb{P} (\text{at some time } t\in[0, T_L], \sigma'|_{B_L} \equiv +1) \geq 1- (1-\epsilon')^{T_L} > 1- \epsilon. \] \end{proof} \begin{lemma}\label{lemmacaptured} If $\rho^+ >0$, then \begin{equation*} \lim_{L \rightarrow \infty}\mathbb{P}(B_l \text{ is captured for some } l \leq L)=1. \end{equation*} \end{lemma} \begin{proof} By Proposition~\ref{schonmann} with $p=\rho^+$, pick a sequence of increasing box sizes $L_i \in \mathbb{N}$ such that $L_1 < L_2 < \ldots$ and \begin{equation*} \mathbb{P} (B_{L_i} \text{ is not entrapped}) < \frac{1}{i^2}. \end{equation*} \noindent By the Borel-Cantelli Lemma, almost surely, all but finitely many boxes $B_{L_i}$ are entrapped. Now the probability that each box $B_{L_i}$ has all corners frozen plus equals \begin{equation*} \mathbb{P} (B_{L_i} \text{ has all corners frozen plus}) = (\rho^+)^{2^d} >0. \end{equation*} \noindent By the Law of Large Numbers, almost surely infinitely many boxes $B_{L_i}$ have this property. Combining these statements, almost surely infinitely many boxes $B_{L_i}$ are captured, which implies the conclusion of the lemma. \end{proof} The remaining lemmas will be used in the proof of Theorem \ref{thm_frozen_+-}. \begin{lemma}\label{lemmaMcaptured} If $\rho^+ > 0$, then \begin{equation*} \lim_{M, L \rightarrow \infty} \mathbb{P} (B_L \text{ is $M$-captured}) =1, \end{equation*} \noindent where $M, L$ tend to infinity with no restriction other than $M \leq L$. \end{lemma} \begin{proof} By Proposition \ref{schonmann}, as in the proof of Lemma \ref{lemmacaptured}, \begin{equation*} \lim_{L \rightarrow \infty} \mathbb{P} (B_L \text{ is entrapped}) =1. \end{equation*} \noindent Now for any fixed $L$ and $M\leq L$, let $A_{L, M}$ denote the event that there exist frozen plus spins within distance $M$ from each of the $2^d$ corners of $B_L$ in every one of the $d$ coordinate directions, as in Definition \ref{Mcaptured}. Thus the event that $B_L$ is $M$-captured is the intersection of $A_{L, M}$ with the event that $B_L$ is entrapped. The probability of the event that any specific collection of $M$ vertices contains no frozen plus spins is $(1-\rho^+)^M$, so \begin{equation*} \mathbb{P} (A_{L, M}) \geq ([1-(1-\rho^+)^M]^d)^{2^d}, \end{equation*} \noindent and this tends to 1 as $M \rightarrow \infty$ (for fixed $d$) uniformly in $L \geq M$. \end{proof} \begin{definition}\label{Mtrimming} Let $B$ be a box of the form $B_L(x)$. We say that $B$ is \textbf{$M$-good} if $B$ is $M$-captured and contains no frozen minus vertices. We define $B[M]$, the \textbf{$M$-trimming} of $B$ as \begin{equation*} B\setminus \left( \bigcup_{i=1}^{2^d} \bar{C}^i (M) \right), \end{equation*} \noindent where each $\bar{C}^i(M)$ is a cube within $B$ containing exactly $M^d$ vertices including the i\textsuperscript{th} corner of $B$ - see Figure \ref{fig:Mtrimming} for the case $d=2$. \end{definition} \begin{figure}[h] \centering \includegraphics{Zdfrozen-3.pdf} \caption{$B[M]$ (grey), the $M$-trimming of an $M$-good box $B$.} \label{fig:Mtrimming} \end{figure} \begin{lemma}\label{lemmaprobMgood} Given $\rho^+ >0$ and $\epsilon >0$, there exist $L<\infty$ and $M<L$ such that for all sufficiently small $\rho^-$ (depending on $d, L, M, \epsilon$ and $\rho^+$), \begin{equation*} \mathbb{P} (B_L \text{ is M-good}) > 1-\epsilon. \end{equation*} \end{lemma} \begin{proof} By Lemma \ref{lemmaMcaptured}, we may choose $L,M$ with $M < L$ so that \begin{equation*} \mathbb{P} (B_L \text{ is } M\text{-captured}) > 1-\frac{\epsilon}{2}. \end{equation*} \noindent We may also pick $\rho^-$ small enough so that the probability that $B_L$ contains any frozen minus vertices is less than $\frac{\epsilon}{2}$. Thus \[ \mathbb{P} (B_L \text{ is not } M\text{-good}) < \epsilon/2 + \epsilon/2 = \epsilon. \] \end{proof} \begin{lemma}\label{lemmaB[M]} Let $M<L$ and $E_{M,L}$ be the event that $B_L$ is $M$-good. Then \[ \mathbb{P}(\text{for any }\sigma'(0), \text{ all vertices in }B_L[M] \text{ fixate to } +1 \mid E_{M,L}) = 1. \] \end{lemma} \begin{proof} If $B_L$ is $M$-good, then by Lemma \ref{lemmabootstrap}, almost surely for some time $t_0$, $\sigma'(t_0)|_{B_L} \equiv +1$. In this case, if $M>0$, $\sigma'(t)|_{B_L}$ need not stay identically $+1$ for $t>t_0$ because vertices near the corners can change from $+1$ to $-1$. But a moment's thought shows that the only vertices near a corner $C^i$ that can change to $-1$ are those in a subset of the cube $\bar{C}^i(M)$ - see Definition \ref{Mtrimming}. This is because the frozen plus vertices $C^{i, j}$ (from Definition \ref{Mcaptured}) protect against the flipping of plus vertices beyond a rectangular parallelipiped contained in $\bar{C}^i(M)$. Note that because $M<L$, every vertex in $B_L[M]$ has at least $d+1$ neighbors in $B_L[M]$. For example, for $d=2$ (see Figure \ref{fig:Mtrimming}), $B_L[M]$ is the union of two rectangles each of width at least three (hence at least two) so that every vertex in a rectangle has at least three neighbors in the rectangle. Thus $B_L [M]$ will have $\sigma'(t)|_{B_L[M]} \equiv +1$ for all $t \geq t_0$. Of course the same arguments apply to any translated box $B_L(x)=B_L +x $ and to $B_{L} [M] (x)= B_{L} [M] + x$. \end{proof} \subsection{Proofs of main results} The first of the two theorems follows easily from Lemmas \ref{lemmabootstrap} and \ref{lemmacaptured}. \begin{proof}[Proof of Theorem \ref{thm_frozen_+d}] If the box $B_L$ has all $2^d$ corners frozen plus, and if at some time $t_0$, \\ $\sigma'(t_0) |_{B_L} \equiv +1$, then $\sigma'(t) |_{B_L} \equiv +1$ for all $t \geq t_0$. This is because after time $t_0$ every vertex in $B_L$ (other than the corners whose spin value is frozen) will have at least $d+1$ plus neighbors, so it won't flip sign. If $B_L$ is also captured, then by Lemma \ref{lemmabootstrap}, with probability one, $\sigma'(t) |_{B_L} \equiv +1$ will occur for some $t$. Now by Lemma \ref{lemmacaptured}, with probability one, $B_L$ will be captured for some $L$ and so $\sigma'(0)$ will fixate to $+1$. The same argument can be translated to $x$ and $B_L(x)$ for any $x\in \mathbb{Z}^d$. \end{proof} \begin{proof}[Proof of Theorem \ref{thm_frozen_+-}] We first introduce an auxiliary graph $G$, with vertex set $\mathbb{Z}^d$, but where $y_1, y_2$ are neighbors (with edge $\{y_1, y_2 \}$) if $\| y_1 - y_2 \|_\infty =1$, so every $y$ has $3^d-1$ neighbors. Pick $M<L$ to be determined later and tile $\mathbb{Z}^d$ with boxes $C_L (y)= B_L ((2L+1)y), ~y\in \mathbb{Z}^d$. For disjoint such boxes the events of being $M$-good are independent, so the collection of $M$-good boxes and $M$-bad (that is, not $M$-good) boxes defines an independent percolation process on a ``renormalized'' copy of $G$, called $\hat G$, by referring to a vertex $y \in \hat G$ as {\bf good} if $C_L(y)$ is $M$-good in $G$ and {\bf bad} otherwise. Note that when two vertices in $\hat G$ are neighbors, this corresponds (for example, for $d=3$), to $C_L(y_1)$ and $C_L(y_2)$ sharing either a face or an edge or just a corner. (The reason for using this notion of connectedness is that a standard (using only nearest neighbor edges) cluster of fixed minus and flipping vertices can extend beyond the standard cluster of $M$-bad (that is, not $M$-good) boxes into the $G$-cluster of $M$-bad boxes and beyond into the $G$-closure of that $G$-cluster.) Let $\mathcal{C}$ denote the cluster of bad vertices containing the origin in this independent site percolation model of sites in $\hat G$, and let $\bar{\mathcal{C}}$ denote its closure (that is, $\mathcal{C}$ unioned with the set of good sites that are $G$-neighbors of sites in $\mathcal{C}$). We will show that the $G$-cluster $\mathcal{C}^\ast$ containing the origin, consisting of (eventually) fixed minus vertices together with flipping vertices of $\sigma'$, satisfies \begin{equation}\label{Cast} \mathbb{P}\left( \mathcal{C}^\ast \subseteq \bigcup_{y\in \bar{\mathcal{C}}} C_L (y) \text{ for every } \sigma'(0) \right)=1. \end{equation} Once Equation~(\ref{Cast}) has been verified, the proof is completed as follows. By standard percolation arguments, there is some $p^\ast >0$ (one can take, for instance, $p^\ast = 1 / (3^d-1)$, since $3^d-1$ is the number of neighbors of any vertex in $G$), such that, if \begin{equation} \label{probbadsite} \mathbb{P} ( y \text{ is bad}) = \mathbb{P}(B_L \text{ is not }M\text{-good})<p^\ast, \end{equation} \noindent then there is almost surely no percolation of bad sites and $\mathbb{E}(|\bar{\mathcal{C}}|) < \infty$. To finish the proof we use Lemma \ref{lemmaprobMgood} to choose $\rho^-$ so small that inequality (\ref{probbadsite}) is valid, and note that by inequality (\ref{Cast}), \begin{equation*} | \mathcal{C}^\ast| \leq |\bar{\mathcal{C}}| (2L+1)^d < \infty. \end{equation*} \end{proof} It remains to prove~(\ref{Cast}), for which we review some old and and introduce some new notation. We refer to vertices $x$ and $y$ in (the original) $\mathbb{Z}^d$ as $G$-neighbors if $\|x-y\|_{\infty}=1$ (and we view them as vertices in the graph $G$). We refer to two boxes $C_L(y_1)$ and $C_L(y_2)$ as $G$-box-neighbors if $y_1$ and $y_2$ are $G$-neighbors in (the renormalized) $\mathbb{Z}^d$ (and we view these vertices in the graph $\hat G$). $G$-clusters and $G$-box-clusters (of certain vertices and boxes) and their boundaries will be used later; their definitions are analogous. Note that there exists a $G$-path connecting a vertex in $C_L(y_1)$ to one in $C_L(y_2)$ and contained within $C_L(y_1) \cup C_L(y_2)$ iff $C_L(y_1)$ and $C_L(y_2)$ are $G$-box-neighbors. A corner region $R(y)$ of a box $C_L(y)$ is a cube entirely within $C_L(y)$ containing exactly $M^d$ vertices including one of the $2^d$ corner vertices of $C_L(y)$. (A vertex is a corner vertex if it has only $d$ neighbors within $C_L(y)$). Any box $C_L(.)$ contains exactly $2^d$ distinct corner regions and their $G$-boundaries are disjoint when $M<L$. We call a box {\bf good} if it is $M$-good. If a box is not good, it is {\bf bad}. We call a vertex bad if it is either in a bad box or in one of the corner regions of a good box; otherwise, it is called a good vertex. Note that by Lemma~\ref{lemmaB[M]}, all good vertices fixate to $+1$. \begin{lemma} \label{GdBxPth} Let $x_1 \in R_1(x)$ and $x_2 \in R_2(x)$ where $R_1(x)$ and $R_2(x)$ are two distinct corner regions of a good $C_L(x)$. Let $\gamma$ be a $G$-path of bad vertices connecting $x_1$ and $x_2$; then $\gamma$ contains a vertex $z$ such that $z \notin C_L(x)$. \end{lemma} \begin{proof} By way of contradiction, suppose such a $G$-path $ \gamma $ exists in $C_L(x)$. After translating $C_L(x)$ to $\{1,2,\dots,2L+1\}^d$, we may assume that $\gamma$ starts at a vertex $\gamma _1 = (a_1, a_2,\ldots, a_d)$ such that for all $i$, $a_i \leq M$. If $\gamma$ ends in a different corner region, denote the last vertex in $\gamma$ by $\gamma_2=(b_1, b_2, \ldots, b_d)$ and note that at least for one $i$, $b_i> M+3$. Without loss in generality, assume $b_1 >M+3$ and note that the $G$-neighbors connecting $\gamma_1$ and $\gamma_2$ differ in any coordinate by at most $1$. Thus there exists a bad vertex $\gamma_3=(c_1, c_2, ...c_d)$ in $\gamma$ with $c_1=M+1$. Recall, however, that in a good box only the corner regions contain bad vertices, so that any site $x=(x_1, x_2, \ldots, x_d)$ with $M+1 \leq x_i \leq M+3$ for at least one $i$, is good. \end{proof} \begin{definition} \label{Adj} For $x\neq y$, two corner regions, $R(x)$ and $R(y)$, of boxes $C_L(x)$ and $C_L(y)$ are called adjacent if there exists $v \in R(x)$ and $v' \in R(y)$ such that $v$ and $v'$ are $G$-neighbors. \end{definition} The proof of the next lemma is straightforward and is left to the reader. \begin{lemma} \label{ADJ} Let $x,y$ be distinct vertices in $\mathbb{Z}^d$. \begin{itemize} \item{}If there is a $G$-path connecting $R(x)$ and $C_L(y)$ and remaining in $C_L(x) \cup C_L(y)$ then $C_L(y)$ has a corner region $R(y)$ adjacent to $R(x)$. \item{} If $R(x)$ is adjacent to $R(y)$ and $R(y)$ is adjacent to $R(z)$ then $R(x)$ is adjacent to $R(z)$ (transitivity of the adjacency relation). \end{itemize} \end{lemma} \begin{lemma} \label{AdjCorPth} Let $C_L(y)$ be a good box and $C_L(x)$ be a bad one. Assume that $C_L(y)$ is a G-box-neighbor of $C_L(x)$. Let $\gamma$ be a $G$-path of bad sites starting in $R(y)$, where $R(y)$ is adjacent to a corner region of $C_L(x)$. Then $\gamma$ can only exit $C_L(y)$ into either \begin{enumerate} \item a bad box that is a $G$-box-neighbor of $C_L(x)$, or \item a corner region, adjacent to $R(y)$, of a good box that is a $G$-box-neighbor of $C_L(x)$. \end{enumerate} \end{lemma} \begin{proof} If the path exits $R(y)$ into a good box $C_L(z)$, then it must (by the reasoning of Lemma~\ref{GdBxPth}) enter into a corner region that is adjacent to $R(y)$ and hence by Lemma~\ref{ADJ}, $C_L(z)$ is a $G$-box-neighbor of both $C_L(y)$ and $C_L(x)$. If it exits into a bad box, $C_L(z')$, then it must enter into an adjacent corner region $R(z')$ or into its $G$-boundary, but in either case, $C_L(z')$ is again a $G$-box-neighbor of both $C_L(y)$ and $C_L(x)$. \end{proof} The next proposition is an immediate consequence of the previous lemma in the setting where $C_L(x)$ is in a $G$-box-cluster of bad boxes. \begin{proposition} There is no $G$-path of bad sites from a $G$-box cluster of bad boxes that exits the union of the $G$-box cluster and its $G$-box-boundary. \end{proposition} \begin{proof}[Proof of Theorem 3.4] \noindent The proof uses the version of Proposition~\ref{schonmann} for Schonmann's {\bf modified basic version} of bootstrap percolation as discussed in Example~3 of~\cite{S}. We will focus on the case $d-1=2$. The case $d-1=1$ is easier and can be handled without the use of bootstrap percolation. We leave it for the reader to check that the proof for the cases with $d-1 \geq 3$ proceed essentially the same as for $d-1=2$. For $d-1=2$, we first partition $\mathbb{Z}^2\times \{0, 1, 2, ...K\}$ into disjoint $2\times 2\times (K+1)$ pillars \\ $\mathcal P_{i,j}= \{ (2i, 2j), (2i+1, 2j), (2i, 2j+1), (2i+1, 2j+1)\}\times\{0, 1, 2, ..., K\}$ for $(i,j) \in Z^2$. If at any time, all vertices in $\mathcal P_{i,j}$ are $+1$, then they stay $+1$ after that since the bottom layer is frozen to $+1$, each site in the top layer has 3 neighbors (within the pillar) out of its 5 total neighbors equal to $+1$, and all other sites have 4 neighbors (within the pillar) out of its 6 total neighbors equal to $+1$. A pillar $\mathcal P_{i,j}$ is fixed $+1$ in this way at time zero, with probability $\theta^{4K}$ and the set of such $(i,j)$ are chosen independently of each other. We can now apply the modified Proposition~\ref{schonmann} as long as we can show that if a pillar $\mathcal P_{i,j}$ (at time $t$) has at least two neighboring pillars -- one in each coordinate direction -- all $+1$ (at time $t$), then almost surely by some random time $t+T$, $P_{i, j}$ will have become all $+1$. This last claim is verified by first arguing, like in the proof of Lemma~\ref{lemmaboostrapweak}, there is strictly positive probability that $\mathcal P_{i,j}$ will be all $+1$ by the time $t+1$, and then proceeding as in the proof of Lemma~\ref{lemmabootstrap}. To explain the first argument, suppose the all $+1$ neighboring pillars are $\mathcal P_{i-1,j}$ and $\mathcal P_{i,j-1}$. Then the vertices in the first layer (above the frozen sites), $(2i, 2j, 1), (2i+1, 2j, 1), (2i, 2j+1, 1), (2i+1, 2j+1, 1)$ can flip to $+1$ (if they are not already $+1$) in that order followed by the sites in layers $2, 3, \ldots, K$ in that order. \end{proof} \bigskip {\bf Acknowledgments.} The authors thank Leo T. Rolla for many fruitful discussions. The research reported in this paper was supported in part by NSF grants DMS-1007524 (S.E. and C.M.N.), DMS-1419230 (M.D.) and OISE-0730136 (S.E., H.K. and C.M.N.). V.S. was supported by ESF-RGLIS network and by Brazilian CNPq grants 308787/2011-0 and 476756/2012-0 and FAPERJ grant E-26/102.878/2012-BBP. \bigskip
\section{Introduction} Separation occurs in flow configurations with abrupt geometry changes or strong adverse pressure gradients. In practical engineering applications, separation is generally associated with low-frequency fluctuations which can have undesirable effects, \eg deterioration of vehicle performance, fatigue of mechanical structures, \etc The control of separated flows is therefore an active research area. Part of the ongoing research work focuses on the laminar regime, where separated flows are steady at low Reynolds number and become unsteady above a threshold value. In this regime, stability theory can help design control strategies by providing insight into the physical phenomena involved in the transition to unsteadiness through, for example, the transient growth of particular initial perturbations, or the bifurcation of unsteady eigenmodes. Examples of separated flows commonly studied as archetypical configurations because of their fundamental interest include bluff bodies (\eg square and circular cylinders), backward-facing steps, bumps, stenotic geometries, and pressure-induced separations over flat plates. Sensitivity analysis uses a variational approach to calculate efficiently the linear sensitivity of some quantity to a modification of the flow or to a given actuation, thus suppressing the need to resort to exhaustive parametric studies. \citet{Hill92AIAA} applied sensitivity analysis to the flow past a cylinder and computed the sensitivity of the most unstable growth rate to passive control by means of a second smaller cylinder, and successfully reproduced most sensitive regions previously identified experimentally by \citet{Stry90}. Since then, sensitivity analysis gained popularity and was applied to evaluate sensitivity to flow modification or to passive control in various flows, both in local and global frameworks. For example, \citet{cor01tcfd} designed a control strategy based on such a variational technique in order to reduce optimal transient growth in boundary layers. This was achieved by computing the sensitivity of an objective function involving energy, which was then iteratively minimized. Such quadratic cost functionals are very often employed in control theory, but sensitivity analysis can be applied to non-quadratic quantities as well. \citet{Bew01} minimized several kinds of cost functionals and successfully relaminarized the turbulent flow in a plane channel using wall transpiration. \citet{Bot03} used a variational approach to compute the sensitivity of eigenvalues to base flow modification in the parallel plane Couette flow, as well as the most destabilising modification. \citet{mar08cyl} studied the sensitivity of the cylinder flow leading eigenvalue to base flow modification and to steady forcing in the bulk and, again, reproduced the regions of \citet{Stry90}. \citet{Mel10} managed to control the first oscillating eigenmode in the compressible flow past a slender axisymmetric body by considering its sensitivity to steady forcing, both in the bulk (with mass, momentum or energy sources) and at the wall (with blowing/suction or heating). Recently, \citet{Bra11} also applied sensitivity analysis to evaluate the effect of steady control on noise amplification (maximal energy amplification under harmonic forcing in steady-state regime) in a globally stable flat-plate boundary layer. In the present study, sensitivity analysis is applied to another quantity of interest in separated flows: the length of the recirculation region $\rl$. Many authors observed that in separated flows the recirculation length increases with $\Rey$ (below the onset of instability): circular cylinder (experimental study by \citet{Tan56}, numerical study by \citet{Gia07}), backward-facing step (experimental study by \citet{Acr68}, numerical study by \citet{bar02}), wall-mounted bump (numerical study by \citet{Mar03}, experimental study by \citet{pas12}), \etc As the recirculation region gets longer, both maximal backward flow and maximal shear increase. From a local stability viewpoint, this tends to destabilise the flow. In addition, since the shear layer elongates, incoming or developing perturbations are amplified over a longer distance while advected downstream, and any region of absolute instability is increased in length too. When the flow becomes unstable and unsteady, as is the case for the cylinder flow above threshold ($\Rey > \Rey_c$), the mean recirculation length decreases \citep{nis78}. This is interpreted as the result of a mean flow correction, and the decrease in the mean value of $\rl$ naturally appears as a characteristic global order parameter of the bifurcation \citep{zie97}. The recirculation length therefore appears as a relevant macroscopic scalar parameter to characterize separated flows. This motivates the design of control strategies which directly target $\rl$, rather than eigenmode growth rates, transient growth, or noise amplification. In other words, we propose control strategies which do not focus on the fate of perturbations but act upon a feature of the base flow itself. We choose to design control configurations based on the steady-state base flow, and consider both subcritical and supercritical Reynolds numbers, $40 \leq \Rey \leq 120$. In the supercritical regime $\Rey > \Rey_c$, the uncontrolled steady-state base flow is linearly unstable and the actual flow observed in experiments or numerical simulations is unsteady; but the sensitivity of the steady-state recirculation length is of interest since reducing $\rl$ might restabilise the flow. The stability of the controlled flow will be assessed systematically to determine when this approach is relevant. This paper is organized as follows. Section~\ref{sec:problem} details the problem formulation and numerical methods. In particular, analytical expressions are derived for sensitivity of recirculation length to base flow modifications and to steady control, both volume forcing and wall blowing/suction. Results are presented in section~\ref{sec:results}: regions sensitive to forcing are identified, and several control configurations which allow to reduce $\rl$ are selected to illustrate the method and to validate the sensitivity analysis against fully non-linear simulations. The linear stability properties of these controlled flows are investigated and discussed in section~\ref{sec:stability}. Conclusions are drawn in section~\ref{sec:conclusion}. \section{Problem formulation and numerical methods} \label{sec:problem} \begin{figure} \vspace{0.4 cm} \centerline{ \begin{overpic}[width=8 cm,tics=10]{fig1.eps} \put(-5, 45) {$\Gamma_{in}$} \put(95, 45) {$\Gamma_{out}$} \put(55, 53.5) {$\Gamma_{lat}$} \put(24.5, 19.5) {$\Gamma_{w}$} \put(101, 26) {$x$} \put(31.5, 59) {$y$} \put(57.5, 24.2) {$\mathbf{x}_r$} \put(45, 33.5) {$\rl$} \put(26, 29) {$R$} \end{overpic} } \caption{ Schematic of the problem geometry and computational domain.} \label{fig:geom_and_domain} \end{figure} The sensitivity of the recirculation length in a two-dimensional incompressible cylinder flow is investigated. A cylinder of radius $R$ is located in a uniform flow. The fluid motion is described by the velocity field $\mathbf{U}=(U,V)^T$ of components $U$ and $V$ in the streamwise and cross-stream directions $x$ and $y$, and the pressure field $P$. The state vector $\mathbf{Q}=(\mathbf{U},P)^T$ is solution of the two-dimensional incompressible Navier--Stokes equations \be \label{eq:NS} \bnabla \bcdot \mathbf{U} = 0, \quad \partial_t \mathbf{U} + \bnabla \mathbf{U} \bcdot \mathbf{U} + \bnabla P - \Rey^{-1} \bnabla^2 \mathbf{U} = \mathbf{F} \ee where $\Rey=U_\infty D/\nu$ is the Reynolds number based on the cylinder diameter $D=2R$, the freestream velocity $U_\infty$ and the fluid kinematic viscosity $\nu$, and $\mathbf{F}$ is a steady volume forcing in the bulk. The following boundary conditions are prescribed: uniform velocity profile $\mathbf{U}_b=(\U_\infty,0)^T$ at the inlet $\Gamma_{in}$, symmetry condition $\partial_y U_b=0, V_b=0$ on lateral boundaries $\Gamma_{lat}$, outflow condition $-P_b \mathbf{n}+\Rey^{-1}\bnabla \mathbf{U}_b\bcdot\mathbf{n}=\mathbf{0}$ on $\Gamma_{out}$, where $\mathbf{n}$ is the normal unit vector oriented outward the domain, blowing/suction $\mathbf{U}_b=\mathbf{U}_w$ on the wall control region $\Gamma_c$, and no-slip condition $\mathbf{U}_b=\mathbf{0}$ on the remaining cylinder wall region $\Gamma_w \setminus \Gamma_c$. In this paper attention is restricted to steady flows $\mathbf{Q}_b(x,y)$ which satisfy: \be \label{eq:NSsteady} \bnabla \bcdot \mathbf{U}_b = 0, \quad \bnabla \mathbf{U}_b \bcdot \mathbf{U}_b + \bnabla P_b - \Rey^{-1} \bnabla^2 \mathbf{U}_b = \mathbf{F}. \ee \subsection{Sensitivity of recirculation length} Assuming the flow is symmetric with respect to the symmetry axis $y=0$, the recirculation length is defined as the distance from the cylinder wall rearmost point $(R,0)$ to the reattachment point $\mathbf{x}_r = (x_r,0)$ as shown in figure~\ref{fig:geom_and_domain}: \be \rl = x_r-R. \label{eq:rl1} \ee The reattachment point is characterized by zero streamwise velocity, $U(x_r,0)=0$, and can therefore be computed with a bisection method on $U_c(x)=U(x,0)$ along the symmetry axis. Throughout this study, only flow modifications and forcings which are symmetric with respect to the symmetry axis will be considered; they result in symmetric flows, thus ensuring that the recirculation length (\ref{eq:rl1}) is well defined. \subsubsection{Sensitivity to base flow modification} Considering a small modification of the base flow $\boldsymbol{\delta} \mathbf{Q}$, the variation of the recirculation length is expressed at first order as \be \delta \rl = (\bnabla_\mathbf{Q} \rl \,|\, \boldsymbol{\delta} \mathbf{Q}) \label{eq:dl} \ee where $\bnabla_\mathbf{Q} \rl = (\bnabla_\mathbf{U} \rl , \bnabla_P \rl)^T$ is the sensitivity to base flow modification, $ (\mathbf{a} \,|\, \mathbf{b}) = \int_{\Omega} \bar\mathbf{a} \bcdot \mathbf{b} \,\mathrm{d}\Omega$ denotes the two-dimensional inner product for real or complex fields, and the overbar stands for complex conjugate. To allow for the calculation of this sensitivity, the recirculation length is rewritten as \be \rl = \int_{R}^{\infty} H \left( - U_c(x) \right) \,\mathrm{d}x = \int_{R}^{\infty} G \left( x \right) \,\mathrm{d}x, \label{eq:rl2b} \ee where $U_c(x)=U(x,0)$ is the streamwise velocity on the symmetry axis and $H$ is the Heaviside step function defined as $H(\alpha)=0$ for $\alpha<0$ and $H(\alpha)=1$ for $\alpha>0$. As illustrated in figure \ref{fig:profiles}, the integrand is equal to 1 in the recirculation region where $U_c(x)<0$, and is equal to 0 downstream, therefore integrating along $x$ from the rear stagnation point gives the recirculation length. \begin{figure} \psfrag{Uc}[r][][1][-90]{$U_c$} \psfrag{H}[r][][1][-90]{$G$} \psfrag{x}[][][1][0]{$x$} \centerline{ \begin{overpic}[height=4.3 cm,tics=10]{fig2a.eps} \put(12.5,57.5) {$(a)$} \put(67,38){\footnotesize $\Rey=20$} \put(76,12) {\footnotesize $\Rey=120$} \end{overpic} \hspace{0.1 cm} \begin{overpic}[height=4.3 cm,tics=10]{fig2b.eps} \put(8,60) {$(b)$} \put(37,37) {$G=1$} \put(74,22) {$G=0$} \put(54.5,16) {$x_r$} \end{overpic} } \caption{ $(a)$ Streamwise velocity along the symmetry axis at $\Rey=20, 40 \ldots 120$. $(b)$~Integrand in the definition of the recirculation length (\ref{eq:rl2b}), illustrated at $\Rey=60$. } \label{fig:profiles} \end{figure} Using the same Lagrangian formalism as \citet{Hill92AIAA}, the recirculation length variation due to a base flow modification $\boldsymbol{\delta} \mathbf{Q} =(\boldsymbol{\delta}\mathbf{U}, \delta P)^T$ is obtained as: \begin{subeqnarray} \label{eq:dlc} \delta \rl & = & \lim_{\epsilon \rightarrow 0} \frac{1}{\epsilon} \left[ \rl(\mathbf{Q}+\epsilon \boldsymbol{\delta} \mathbf{Q})-\rl(\mathbf{Q}) \right] \slabel{eq:dlc0} \\ & = & \lim_{\epsilon \rightarrow 0} \frac{1}{\epsilon} \int_{R}^{\infty} \left[ H \left( - U_c(x)-\epsilon \delta U_c(x) \right) - H \left( -U_c(x) \right) \right] \,\mathrm{d}x \slabel{eq:dlc1} \\ & = & \int_{R}^{\infty} -\left.\frac{ \mathrm{d}H }{ \mathrm{d}U }\right|_{U=-U_c(x)} \,\delta U_c(x) \,\mathrm{d}x \slabel{eq:dlc2} \\ & = & \int_{R}^{\infty} \left( \frac{\mathrm{d} U_c}{\mathrm{d} x}(x) \right)^{-1} \frac{ \mathrm{d}G }{ \mathrm{d}x }(x) \,\delta U_c(x) \,\mathrm{d}x \slabel{eq:dlc3} \\ & = & \int_{R}^{\infty} - \left( \frac{\mathrm{d} U_c}{\mathrm{d} x}(x) \right)^{-1} \delta(x-x_r) \,\delta U_c(x) \,\mathrm{d}x \slabel{eq:dlc4} \\ & = & - \frac{ \delta U_c(x_r) } { \mathrm{d}_x U_c|_{x=x_r}}, \slabel{eq:dlc5} \end{subeqnarray} where (\ref{eq:dlc3}) comes from the differentiation of $G(x)=H(-U_c(x))$ using the chain rule, $\displaystyle \frac{\mathrm{d} G}{\mathrm{d} x}(x) $ $\displaystyle= -\left.\frac{\mathrm{d} H}{\mathrm{d} U}\right|_{U=-U_c(x)} \frac{\mathrm{d} U_c}{\mathrm{d} x}(x)$, and (\ref{eq:dlc4}) is the result of $\displaystyle \frac{\mathrm{d} G}{\mathrm{d} x}(x)=-\delta(x-x_r)$ with $\delta(x)$ the Dirac delta function, since $G$ jumps from 1 to 0 at $x=x_r$. If, for example, the streamwise velocity at the original reattachment point increases, then the recirculation region is shortened: this is understood physically as $U_c(x_r)$ becoming positive and the reattachment point moving upstream, while mathematically $\delta U_c(x_r)>0$ and (\ref{eq:dlc}) yield $\delta \rl <0$ (because $\mathrm{d}_x U_c|_{x=x_r}>0$, see figure \ref{fig:profiles}$(a)$). The sensitivity to base flow modification is identified as: \be \bnabla_\mathbf{U} \rl = -\frac{1}{\mathrm{d}_x U_c|_{x=x_r}} \left( \begin{array}{cc} \delta(x_r,0) \\ 0 \end{array} \right), \quad \nabla_P \rl = 0, \label{eq:sens_open} \ee where $\delta(x,y)$ is the two-dimensional Dirac delta function, such that the inner product (\ref{eq:dl}) between the two fields $\bnabla_\mathbf{Q} \rl$ and $\boldsymbol{\delta} \mathbf{Q}$ is indeed $\delta \rl$ as expressed by (\ref{eq:dlc}). In wall-bounded flows, where reattachment occurs at the wall (for example behind a backward-facing step or a bump, or on a flat plate with adverse pressure gradient), the reattachment point is not characterized by zero streamwise velocity $U(x_r,0)=0$, but instead by zero wall shear stress, i.e. $\partial_y U|_{x=x_r,y=0}=0$ with the wall assumed horizontal and located at $y=0$ for the sake of simplicity. In this case the sensitivity reads: \be \bnabla_\mathbf{U} \rl = -\frac{1}{\partial_{xy} U|_{x=x_r,y=0}} \left( \begin{array}{cc} \delta(x_r,0) \partial_y \\ 0 \end{array} \right), \quad \nabla_P \rl = 0. \label{eq:sens_wallbounded} \ee The sensitivity field $\bnabla_\mathbf{Q} \rl$ in (\ref{eq:sens_open})-(\ref{eq:sens_wallbounded}) is valid for any arbitrary base flow modification $\boldsymbol{\delta}\mathbf{U}$. As noted by \citet{Bra11}, it is possible to derive a restricted sensitivity field for divergence-free base flow modifications. In the case of the cylinder flow, where (\ref{eq:sens_open}) results in a localized Dirac delta function at the reattachment point in the $x$ direction only, this restricted sensitivity field appears to present a dipolar structure. \subsubsection{Sensitivity to forcing} Now the sensitivity of the recirculation length to steady forcing is investigated. One considers a small-amplitude forcing: volume force in the bulk $\boldsymbol{\delta} \mathbf{F}(x,y)$, or blowing/suction $\boldsymbol{\delta} \mathbf{U}_w$ on part of the cylinder wall. The recirculation length variation at first order is \be \delta \rl = (\bnabla_\mathbf{F} \rl \,|\, \boldsymbol{\delta} \mathbf{F}) + \langle \bnabla_{\mathbf{U}_w} \rl \,|\, \boldsymbol{\delta} \mathbf{U}_w \rangle \label{eq:RecircLengthVarForce} \ee where $\langle \mathbf{a} \,|\, \mathbf{b} \rangle = \int_{\Gamma_c} \bar\mathbf{a} \bcdot \mathbf{b} \,\mathrm{d}\Gamma$ denotes the one-dimensional inner product on the control boundary. The same Lagrangian formalism as in the previous section yields the sensitivities \be \bnabla_\mathbf{F} \rl = \mathbf{U}^\dag, \quad \bnabla_{\mathbf{U}_w} \rl = \mathbf{U}^\dag_w = -\Pa \mathbf{n} - \Rey^{-1} \bnabla \mathbf{U}^\dag \bcdot \mathbf{n}, \ee where the so-called adjoint base flow $\mathbf{Q}^\dag=(\mathbf{U}^\dag,\Pa)^T$ is solution of the linear, non-homogeneous system of equations \be \bnabla \bcdot \mathbf{U}^\dag =0, \quad -\bnabla \mathbf{U}^\dag \bcdot \mathbf{U}_b + \bnabla \mathbf{U}_b^T \bcdot \mathbf{U}^\dag - \bnabla \Pa - \Rey^{-1} \bnabla^2 \mathbf{U}^\dag = \bnabla_\mathbf{U} \rl, \label{eq:adjBF-lc} \ee with the boundary conditions $\mathbf{U}^\dag=\mathbf{0}$ on $\Gamma_{in} \cup \Gamma_w \cup \Gamma_c$, symmetry condition $\partial_y \Ua=0, \Va=0$ on $\Gamma_{lat}$, and $\Pa\mathbf{n}+\Rey^{-1}\bnabla\mathbf{U}^\dag\bcdot\mathbf{n}+\mathbf{U}^\dag(\mathbf{U}_b\bcdot\mathbf{n})=\mathbf{0}$ on $\Gamma_{out}$. \subsubsection{Effect of a small control cylinder} \label{sec:smallctrlcyl} It is of practical interest to study the effect of a particular kind of passive control on recirculation length and eigenvalues, namely a small control cylinder of diameter $d \ll D$ similar to the one used by \citet{Stry90} to suppress vortex shedding in a limited range of Reynolds number above instability threshold. The effect on the base flow of a small control cylinder located at $\mathbf{x}_c=(x_c,y_c)$ is modelled as a steady volume force of same amplitude as the drag force acting on this control cylinder, and of opposite direction: \be \boldsymbol{\delta} \mathbf{F}(x,y) = -\frac{1}{2} d C_d(x,y) ||\mathbf{U}_b(x,y)|| \mathbf{U}_b(x,y) \delta(x-x_c,y-y_c) \label{eq:SmallCylForce} \ee where $C_d$ is the drag coefficient of the control cylinder and depends on the local Reynolds number $\Rey_d(x,y)=||\mathbf{U}_b(x,y)|| d/\nu$. Finally, variations of the recirculation length and eigenvalue (see section \ref{sec:stab}) are calculated from $\delta \rl = (\bnabla_\mathbf{F} \rl \,|\, \boldsymbol{\delta} \mathbf{F})$ and $\delta \sigma = (\bnabla_\mathbf{F} \sigma \,|\, \boldsymbol{\delta} \mathbf{F})$: \begin{eqnarray} \delta \rl(x_c,y_c) &=& -\frac{1}{2} d C_d(x_c,y_c) ||\mathbf{U}_b(x_c,y_c)|| \bnabla_\mathbf{F} \rl(x_c,y_c) \bcdot \mathbf{U}_b(x_c,y_c), \label{eq:SmallCylEffectRL} \\ \delta \sigma(x_c,y_c) &=& -\frac{1}{2} d C_d(x_c,y_c) ||\mathbf{U}_b(x_c,y_c)|| \bnabla_\mathbf{F} \sigma (x_c,y_c) \bcdot \mathbf{U}_b(x_c,y_c). \label{eq:SmallCylEffectEV} \end{eqnarray} For a diameter $d=D/10$ and for the set of Reynolds numbers $\Rey$ and locations $(x_c,y_c)$ chosen hereafter, the Reynolds number of the control cylinder falls in the range $1 \leq \Rey_d \leq 15$. The expression of \cite{Hill92AIAA} has been generalized in this range according to $C_d(\Rey_d) = a + b \Rey_d^c$, based on a set of experimental data from \citet{fin53} and \citet{tri59} and from numerical results obtained by the authors, yielding $a=0.8558, b= 10.05, c=-0.7004$. \subsection{Linear stability} \label{sec:stab} Writing the flow as the superposition of a steady base flow and time-dependent small perturbations, $\mathbf{Q}(x,y,t) = \mathbf{Q}_b(x,y) + \mathbf{q}'(x,y,t)$, linearising the Navier--Stokes equations~(\ref{eq:NS}) and using the normal mode expansion $\mathbf{q}'(x,y,t) = \mathbf{q}(x,y) e^{\sigma t}$, with $\sigma = \grow + i {\sigma_i}$, the following system of equations is obtained: \be \label{eq:evp} \bnabla \bcdot \uu = 0, \quad \sigma \uu + \bnabla \uu \bcdot \mathbf{U}_b + \bnabla \mathbf{U}_b \bcdot \uu + \bnabla p - \Rey^{-1} \bnabla^2 \uu = \mathbf{0}, \ee together with the following boundary conditions: $\uu=\mathbf{0}$ on $\Gamma_{in} \cup \Gamma_w \cup \Gamma_c$, symmetry condition $\partial_y u=0, v=0$ on $\Gamma_{lat}$, and outflow condition $- p \mathbf{n}+\Rey^{-1}\bnabla \uu \bcdot \mathbf{n}=\mathbf{0}$ on $\Gamma_{out}$. Solving this generalized eigenvalue problem yields global modes $\mathbf{q}$ and associated growth rate $\grow$ and pulsation ${\sigma_i}$. The sensitivity of an eigenvalue to base flow modification, defined by $\delta \sigma = (\bnabla_\mathbf{U} \sigma \,|\, \boldsymbol{\delta} \mathbf{U})$, can be computed as \be \bnabla_{\mathbf{U}} \sigma = -\bnabla \bar\uu^T \bcdot \uua + \bnabla \uua \bcdot \bar\uu, \ee where $\mathbf{q}^\dag=(\uua,p^\dag)^T$ is the adjoint mode associated with $\sigma$. The sensitivity to steady forcing, defined by $\delta \sigma = (\bnabla_\mathbf{F} \sigma \,|\, \boldsymbol{\delta} \mathbf{F}) + \langle \bnabla_{\mathbf{U}_w} \sigma \,|\, \boldsymbol{\delta} \mathbf{U}_w \rangle$, can be computed as \be \bnabla_\mathbf{F} \sigma = \mathbf{U}^\dag, \quad \bnabla_{\mathbf{U}_w} \sigma = \mathbf{U}^\dag_w = -\Pa \mathbf{n} - \Rey^{-1} \bnabla \mathbf{U}^\dag \bcdot \mathbf{n} \ee where this time the adjoint base flow $(\mathbf{U}^\dag,\Pa)^T$ is solution of the linear system \be \bnabla \bcdot \mathbf{U}^\dag =0, \quad -\bnabla \mathbf{U}^\dag \bcdot \mathbf{U}_b + \bnabla \mathbf{U}_b^T \bcdot \mathbf{U}^\dag - \bnabla \Pa - \Rey^{-1} \bnabla^2 \mathbf{U}^\dag = \bnabla_\mathbf{U} \sigma, \label{eq:adjBF-om} \ee with boundary conditions $\mathbf{U}^\dag=\mathbf{0}$ on $\Gamma_{in} \cup \Gamma_w \cup \Gamma_c$, $\partial_y \Ua=0, \Va=0$ on $\Gamma_{lat}$, and $\Pa \mathbf{n}+\Rey^{-1}\bnabla \mathbf{U}^\dag \bcdot \mathbf{n} + \mathbf{U}^\dag(\mathbf{U}_b\bcdot\mathbf{n}) + \uua(\bar\uu \bcdot \mathbf{n})=\mathbf{0}$ on $\Gamma_{out}$. It is possible to relate the sensitivity of a given eigenvalue and the individual sensitivities of its growth rate and pulsation, $\delta \sigma_{r,i} = (\bnabla_{\mathbf{U}} \sigma_{r,i} \,|\, \boldsymbol{\delta}\mathbf{U})$ according to $\bnabla_{\mathbf{U}} \grow = \Real\{\bnabla_{\mathbf{U}} \sigma\}$ and $\bnabla_{\mathbf{U}} {\sigma_i} = -\Imag\{\bnabla_{\mathbf{U}} \sigma\}$. The same relations hold for sensitivity to forcing. \subsection{Numerical method} \label{sec:num} All calculations are performed using the finite element software \textit{FreeFem++} to generate a two-dimensional triangulation in the computation domain $\Omega$ shown in figure~\ref{fig:geom_and_domain}, of dimensions $-50 \leq x \leq 175$, $-30 \leq y \leq 30$, with the center of cylinder located at $x=0, y=0$. Bold lines indicate boundaries. The mesh density increases from the outer boundaries towards the cylinder wall, in successive regions indicated by dash-dotted lines in figure~\ref{fig:geom_and_domain}. The resulting mesh has 246083 triangular elements. Variational formulations associated to the equations to be solved are spatially discretized using P2 and P1 Taylor-Hood elements for velocity and pressure respectively. Base flows are computed using an iterative Newton method to solve equations~(\ref{eq:NSsteady}), convergence being reached when the residual is smaller than $10^{-12}$ in $L^2$ norm. The eigenvalue problem~(\ref{eq:evp}) is solved using an implicitly restarted Arnoldi method. Adjoint base flows involved in the calculation of recirculation length sensitivity and eigenvalue sensitivity are obtained by inverting the simple linear systems (\ref{eq:adjBF-lc}) and (\ref{eq:adjBF-om}). Convergence was checked by calculating steady-state base flows with different meshes. Reducing the number of elements by 21\%, the recirculation length varied by 0.2\% or less over the range of Reynolds numbers $30 \leq \Rey \leq 120$. Values of drag coefficient and recirculation length over this same $\Rey$ range are given in table~\ref{tab:valid} together with results from the literature. Compared to \citet{Gia07}, the maximum relative difference on $\rl$ was 2\%, of the same order as the values they report, while $C_D$ differed by less than 1.7\% from values computed by \citet{hen95}. From linear stability calculations, the onset of instability characterized by $\grow=0$ was found to be $\Rey_c = 46.6$, in good agreement with the values reported in the literature. Also, the frequency ${\sigma_i}$ of the most unstable global mode differed by less than 1.2\% from results of \citet{Gia07}. \begin{table} \small \begin{center} \def~{\hphantom{0}} \begin{tabular}[]{l ccccc c ccccc c c} & \multicolumn{5}{c}{$C_D$} & & \multicolumn{5}{c}{$\rl$} & $\Rey_c$\\ & $20$ & $40$ & $60$ & $100$ & $120$ & & $20$ & $40$ & $60$ & $100$ & $120$ \\ \cline{2-6} \cline{8-12} \cline{14-14} \citet{hen95} & 2.06 & 1.54 & 1.31 & 1.08 & 1.01 & &&&&& & \\ \citet{zie97} &&&&& & & 0.94 & 2.28 & 3.62 & 6.30 & & \\ \citet{Gia07} & 2.05 & 1.54 & & & & & 0.92 & 2.24 & (3.6) & (6.2) & (7.5) & 46.7 \\ \citet{sip07} &&&&& & &&&&& & 46.6 \\ \citet{mar08cyl} &&&&& & &&&&& & 46.8 \\ Present study & 2.04 & 1.52 & 1.30 & 1.07 & 1.00 & & 0.92 & 2.25 & 3.57 & 6.14 & 7.42 & 46.6 \\ \end{tabular} \caption{\small Drag coefficient $C_D$ and recirculation length $\rl$ for different values of $\Rey$, and critical Reynolds number $\Rey_c$. Bracketed numbers are estimated from a figure.} \label{tab:valid} \end{center} \end{table} \section{Results} \label{sec:results} In this section we consider subcritical and supercritical Reynolds numbers, $40 \leq \Rey \leq 120$, and focus on the steady-state recirculation length. Its sensitivity to flow modification and to control is presented in section~\ref{sec:sensitivity}, and examples of control configurations which reduce $\rl$ are detailed in section~\ref{sec:ctrl}. Stability properties are discussed later in section~\ref{sec:stability}. \subsection{Sensitivity of recirculation length} \label{sec:sensitivity} Figure~\ref{fig:sensit_l_F}$(a)$ shows the sensitivity of recirculation length to bulk forcing in the streamwise direction, $\nabla_{F_x} \rl = \bnabla_{\mathbf{F}} \rl \bcdot \mathbf{e}_x$. By construction, sensitivity analysis predicts that forcing has a large effect on $\rl$ in regions where sensitivity is large. To be more specific, $\rl$ can be reduced by forcing along $\mathbf{e}_x$ in regions of negative sensitivity $\nabla_{F_x} \rl<0$: in the recirculation region (in particular close to the reattachment point), and near the sides of the cylinder just upstream of the separation points; $\rl$ can also be reduced by forcing along $-\mathbf{e}_x$ in regions of positive sensitivity $\nabla_{F_x} \rl>0$: at the outer sides of the recirculation region (in particular close to the reattachment point). \begin{figure} \vspace{0.2 cm} \centerline{ \begin{overpic}[height=11 cm,tics=10]{fig3-new.eps} \put(-2.5,80) {$(a)$} \put(49.,80) {$(b)$} \end{overpic} \vspace{0.2 cm} } \caption{ (a) Normalized sensitivity of recirculation length to bulk forcing in the streamwise direction, $\nabla_{F_x} \rl/\rl$. (b) Normalized effect of a small control cylinder of diameter $d=0.10D$ on recirculation length, $\delta \rl/\rl$. From top to bottom: $\Rey=40, 60, 80, 100, 120$. The dashed line is the steady-state base flow separatrix. Black circles show the locations of control cylinders for configuration B discussed in section~\ref{sec:ctrl}. } \label{fig:sensit_l_F} \end{figure} \begin{figure} \def\thisfigytop{25} \def\thisfigybot{-3} \vspace{0.9cm} \centerline{ \begin{overpic}[width=10.5cm, tics=10]{fig4.eps} \put( 5, \thisfigytop) {\footnotesize $5.52$} \put(25.75, \thisfigytop) {\footnotesize $6.57$} \put(46.50, \thisfigytop) {\footnotesize $7.56$} \put(67.25, \thisfigytop) {\footnotesize $8.50$} \put(88, \thisfigytop) {\footnotesize $9.40$} % \put(2, \thisfigybot) {\footnotesize $\Rey=40$} \put(22.75, \thisfigybot) {\footnotesize $\Rey=60$} \put(43.50, \thisfigybot) {\footnotesize $\Rey=80$} \put(64.25, \thisfigybot) {\footnotesize $\Rey=100$} \put(85, \thisfigybot) {\footnotesize $\Rey=120$} \end{overpic} } \vspace{0.4cm} \caption{ Sensitivity of recirculation length to wall actuation $\bnabla_{\mathbf{U}_w} \rl$. Flow is from left to right. Numbers correspond to the $L^\infty$ norm of $\bnabla_{\mathbf{U}_w} \rl/\rl$. } \label{fig:sensit_l_Uw-ntxy} \end{figure} \begin{figure} \vspace{0.3 cm} \centerline{ \begin{overpic}[height=3.5 cm,tics=10]{fig5.eps} \put(45, 37) {\footnotesize $x$} \put(23, 59) {\footnotesize $y$} \put(85, 54) {\footnotesize $\theta$} \put(-8, 90) {\footnotesize $\delta V_w$} \end{overpic} } \caption{ Sketch of control configuration W. } \label{fig:sketch_wall_forcing} \end{figure} As mentioned in section~\ref{sec:smallctrlcyl}, it is convenient in an experiment to use a small, secondary cylinder as a simple passive control device which produces a force which aligned with the local flow direction and depending non-linearly on the local velocity, as given by (\ref{eq:SmallCylForce}). The effect $\delta \rl$ of such a control cylinder was computed from $\bnabla_{\mathbf{F}} \rl$ according to (\ref{eq:SmallCylEffectRL}). Results for a control cylinder of diameter $d=0.10D$ are shown in figure~\ref{fig:sensit_l_F}$(b)$. The recirculation length increases ($\delta\rl>0$) when a control cylinder is located on the sides of the main cylinder upstream of the separation point , or in the shear layers; it decreases ($\delta\rl<0$) when a control cylinder is located farther away on the sides of the shear layers. The minimal value of $\delta \rl/\rl$ becomes less negative as $\Rey$ increases, meaning that a control cylinder of constant diameter becomes gradually less effective at reducing $\rl$. Interestingly, regions associated with $\rl$ reduction correspond qualitatively well to regions where vortex shedding was suppressed in the experiments of \citet{Stry90} (figure 20 therein) for the same diameter ratio. This point is further discussed in section~\ref{sec:stability}. Figure~\ref{fig:sensit_l_Uw-ntxy} shows the sensitivity of recirculation length to steady wall actuation, $\bnabla_{\mathbf{U}_w} \rl$. Since the variation of $\rl$ is given by the inner product $\delta \sigma = \langle \bnabla_{\mathbf{U}_w} \sigma \,|\, \boldsymbol{\delta} \mathbf{U}_w \rangle$, wall control oriented along the arrows increases $\rl$. In particular, the recirculation length is increased by wall blowing where arrows point towards the fluid domain, and by wall suction where arrows point inside the cylinder. The numbers above each plot correspond to the $L^\infty$ norm of the rescaled sensitivity field $\bnabla_{\mathbf{U}_w} \rl/\rl$. This norm increases (roughly linearly) with $\Rey$, indicating that the relative control authority is increasing. The shape of the sensitivity field does not vary much with $\Rey$, and reveals that the most efficient way to reduce $\rl$ is to use wall suction at the top and bottom sides of the cylinder, in a direction close to wall normal. \subsection{Control of the recirculation length} \label{sec:ctrl} The sensitivity fields obtained in the previous section can be used to control recirculation length. To illustrate this process, two control configurations predicted to reduce $\rl$ are tested at two representative Reynolds numbers $\Rey=60$ and 120, and for different control amplitudes: \begin{description} \item - configuration B (``bulk''; sketched in figure~\ref{fig:sensit_l_F}): volume forcing with two control cylinders located symmetrically close to the point where $\rl$ reduction is predicted to be maximal, at $\mathbf{x}^* = (x^*,\pm y^*) = (1.2,\pm 1.15)$ at $\Rey=60$ and $\mathbf{x}^* = (2.4,\pm 1.3)$ at $\Rey=120$; \item - configuration W (``wall''; sketched in figure~\ref{fig:sketch_wall_forcing}): vertical wall suction at the top and bottom sides, $\pi/3 \leq |\theta| \leq 2\pi/3$, with velocity $\delta V_w$. \end{description} \begin{figure} \psfrag{dU/dF}[t][][1][0]{$\delta F,\,|\delta V_w|$} \psfrag{F} [t][][1][0]{$\delta F$} \psfrag{lc}[br][l][1][-90]{$\rl\,\,\,$} \psfrag{llc}[b][][1][-90]{} \psfrag{d=0.05D} [][][1][0]{\footnotesize $d=0.05D$} \psfrag{d=0.1D} [][][1][0]{\footnotesize $d=0.10D$} \centerline{ \begin{overpic}[height=5 cm,tics=10]{fig6a.eps} \put(-5,75) {$(a)$} \put(25,35) {\footnotesize \textcolor{red}{W}} \put(55,40) {\footnotesize \textcolor{blue}{B}} \end{overpic} \hspace{0.4 cm} \begin{overpic}[height=5 cm,tics=10]{fig6b.eps} \put(-5,75) {$(b)$} \put(17,35) {\footnotesize \textcolor{red}{W}} \put(65,45) {\footnotesize \textcolor{blue}{B}} \end{overpic} } \caption{ Variation of recirculation length with control amplitude ($\delta F$ in configuration B, $|\delta V_w|$ in configuration W). $(a)$ $\Rey=60$, $(b)$ $\Rey=120$. Dashed lines show predictions from sensitivity analysis, solid lines are non-linear results. Symbols correspond to control cylinders of diameter $d=0.05 D$ ($\blacklozenge$) and $d=0.10 D$ ($\scriptstyle\blacksquare$). } \label{fig:sensit_dl_F} \end{figure} \begin{figure} \vspace{0.7cm} \hspace{0.2cm} \centerline{ \begin{overpic}[width=13 cm,tics=10]{fig7-new.eps} \put(-3,42) {$(a)$} \put(47,42) {$(b)$} \put(-3,17) {$(c)$} \put(47,17) {$(d)$} \put(31,38) {\small {Uncontrolled}} \put(33.5,31) {\small {Controlled}} \put(31,13.5) {\small {Uncontrolled}} \put(33.5,6.5) {\small {Controlled}} \put(79.5,38) {\small {Uncontrolled}} \put(82,31) {\small {Controlled}} \put(79.5,13.5) {\small {Uncontrolled}} \put(82,6.5) {\small {Controlled}} \end{overpic} } \caption{ Streamwise velocity of uncontrolled and controlled steady-state base flow. Left: $\Rey=60$; right: $\Rey=120$. $(a)$-$(b)$ configuration B: two control cylinders of diameter $d=0.10 D$ located at $(x^{*},\pm y^{*})$; $(c)$-$(d)$ configuration W: suction of amplitude $\delta V_w=0.05$ and $0.03$ respectively. } \label{fig:modifiedBF} \end{figure} Figure~\ref{fig:sensit_dl_F} shows the recirculation length variation $\delta \rl$ for these configurations. In addition to predictions from sensitivity analysis (dashed lines), this figure shows non-linear results (solid lines) obtained by computing the actual controlled flow and its recirculation length for each configuration and each amplitude. In configuration B, the effect of control cylinders is modelled by the volume force (\ref{eq:SmallCylForce}); in configuration W, wall actuation is implemented as a velocity boundary condition with uniform profile. As expected, $\rl$ is reduced, and the agreement between sensitivity analysis and non-linear results is excellent, with slopes matching at zero-amplitude, whereas non-linear effects appear for larger amplitudes. Non-linear simulations indicate that at $\Rey=60$, controlling with two cylinders of diameter $d=0.10 D$ reduces $\rl$ by more than 65\%; a reduction of about 50\% is achieved with control cylinders of diameter $0.05D$, and with wall suction of intensity $\delta V_w = 0.12$. At $\Rey=120$, control cylinders of diameter $d=0.10 D$ reduce $\rl$ by more than 40\%; a reduction of about 30\% is achieved with control cylinders of diameter $0.05D$, and with wall suction of intensity $\delta V_w = 0.04$. Figure~\ref{fig:modifiedBF} shows examples of controlled flows, illustrating how the recirculation region is shortened. In configuration B, control cylinders of diameter $d=0.10 D$ located at $(x^*,\pm y^*)$ make the flow deviate and accelerate. As a result, the streamwise velocity increases between the two control cylinders in a long region extending far downstream, and the reattachment point moves upstream. In configuration W, fluid is sucked at the cylinder sides, which brings fluid with high streamwise velocity closer to the wall region, which in turn makes the recirculation region shorter. \section{Effect on linear stability} \label{sec:stability} In the previous section, sensitivity analysis was performed on the steady-state base flow. It provided information on the sensitivity of the recirculation length and allowed to design efficient control strategies. These results are relevant only if the controlled flow is stable. Indeed, when increasing the Reynolds number above its critical value $\Rey_c$, the cylinder flow becomes linearly unstable, with a Hopf bifurcation leading to unsteady vortex shedding; one should therefore investigate whether the flow is stabilised by the control. In the subcritical regime $\Rey < \Rey_c$, one should also check that the flow is not destabilised by the control. At $\Rey=60$, one pair of complex conjugate eigenvalues (``mode 1'') associated with the von K\'arm\'an street is unstable. Figure~\ref{fig:sensit_domi_F}$(a)$ shows the variation of the leading growth rate $\sigma_{1,r}$ with control amplitude. Both configurations B and W have a stabilising effect on the leading global mode. Full restabilisation is achieved with two control cylinders of diameter $d \simeq 0.04 D$, or with wall suction of intensity $\delta V_w \simeq 0.08$. These control configurations do not destabilise other eigenmodes for any of the amplitudes tested. \begin{figure} \vspace{0.3 cm} \psfrag{omi}[c][t][1][-90]{$\grow$} \psfrag{growth rate}[c][t][1][-90]{} \psfrag{F} [t][][1][0]{$\delta F$} \psfrag{dU/dF}[t][][1][0]{$\delta F, \, |\delta V_w|$} \centerline{ \begin{overpic}[height=4.7 cm,tics=10]{fig8a.eps} \put(-2,70) {$(a)$} \put(25,67) {\footnotesize Mode 1} \put(42,33) {\footnotesize \textcolor{red}{W}} \put(73,33) {\footnotesize \textcolor{blue}{B}} \end{overpic} \hspace{0.3 cm} \begin{overpic}[height=4.7 cm,tics=10]{fig8b.eps} \put(-2,71) {$(b)$} \put(16,60) {\footnotesize Mode 1} \put(16,33) {\footnotesize Mode 2} \put(46,55) {\footnotesize \textcolor{red}{W}} \put(82,55) {\footnotesize \textcolor{blue}{B}} \put(22,15) {\footnotesize \textcolor{red}{W}} \put(82,15) {\footnotesize \textcolor{blue}{B}} \end{overpic} } \caption{ Variation of leading growth rates with control amplitude. $(a)$ $\Rey=60$, $(b)$~$\Rey=120$. Same notations as in figure~\ref{fig:sensit_dl_F}. } \label{fig:sensit_domi_F} \end{figure} \begin{figure} \def62{62} \vspace{0.2 cm} \hspace{1.7 cm} \centerline{ \begin{overpic}[width=12 cm,tics=10]{fig9-new.eps} \put(62, 92){\small $\delta\rl/\rl$} \put(62, 83){\small $\delta\sigma_{1,r} / |\sigma_{1,r}|$} \end{overpic} } \caption{ Normalized effect of a small control cylinder ($d=0.10D$) on the recirculation length, $\delta\rl/\rl$ (upper half of each panel), and on the most unstable growth rate, $\delta \sigma_{1,r} / |\sigma_{1,r}|$ (lower half). From top to bottom: $\Rey=40, 60, 80, 100, 120$. The stabilising contour $\delta \sigma_{1,r} / \sigma_{1,r}=-1/2$ for two symmetric control cylinders is reported on the upper half plots for $\Rey > \Rey_c$. } \label{fig:sensit_omi_F} \end{figure} A second pair of eigenvalues becomes unstable (``mode 2'') at $\Rey \simeq 110$ \citep{ver11}. Figure~\ref{fig:sensit_domi_F}$(b)$ shows the variation of $\sigma_{1,r}$ and $\sigma_{2,r}$ with control amplitude at $\Rey=120$. Bulk forcing has a stabilising effect on both unstable modes, but does not achieve full restabilisation. The straightforward strategy which allowed to stabilise the flow at $\Rey=60$, namely placing control cylinders where the sensitivity analysis predicts they have the largest reducing effect on $\rl$, is therefore not successful at $\Rey=120$. Wall forcing has a stabilising effect on mode 2 but a destabilising effect on mode 1 for reasonable control amplitude, and again the flow is not restabilised. To investigate why control configurations which shorten the recirculation region also have a stabilising effect at $\Rey=60$, but not at $\Rey=120$, it is interesting to consider the sensitivity of the leading eigenvalue. Figure~\ref{fig:sensit_omi_F} compares the effect of a small control cylinder ($d=0.10D$) on $\rl$, calculated from (\ref{eq:SmallCylEffectRL}) and already shown in figure~\ref{fig:sensit_l_F}$(b)$, and its effect on $\sigma_{1,r}$, calculated from (\ref{eq:SmallCylEffectEV}). At $\Rey=40$ and 60, $\delta\rl$ and $\delta\sigma_{1,r}$ have very similar spatial structures. This means that a small control cylinder, when located where it reduces the recirculation length, almost always has a stabilising effect on the most unstable global mode. However, this similarity gradually disappears as $\Rey$ increases. Regions where $\delta\rl$ and $\delta\sigma_{1,r}$ have opposite signs grow in size, and at $\Rey=120$ they extend along the whole shear layers, both inside and outside the recirculation region. To ease comparison, the contour where two symmetric control cylinders ($d=0.10D$) render mode 1 just neutrally stable (i.e. where $\delta \sigma_{1,r} = - \sigma_{1,r}/2$) is reported on the map of $\delta\rl$. At $\Rey=60$ this stabilising region overlaps with the region of recirculation length reduction, but as $\Rey$ increases it moves upstream towards the region of recirculation length increase. In other words, control cylinders located where they reduce $\rl$ are efficient in stabilising mode 1 at low $\Rey$, but gradually lose this ability at higher $\Rey$. In the latter regime, one may wonder if increasing the recirculation length is not a better way to stabilise the flow. It must be pointed out that the stabilising region shrinks as $\Rey$ increases anyway, consistent with observations from \citet{Stry90} and \citet{mar08cyl}. \begin{figure} \psfrag{x} [ ][ ][1][ 0]{$x$} \psfrag{y} [r][l][1][-90]{$y$} \psfrag{dlc/lc}[][][1][-90]{$\displaystyle \frac{\delta\rl}{\rl} \quad$} \psfrag{dsigr/sigr}[][t][1][-90]{$\displaystyle \frac{\delta\sigma_{1,r}}{\sigma_{1,r}} \quad$} \psfrag{re} [t][][1][0]{$\Rey$} % \centerline{ \begin{overpic}[width=10cm,tics=10]{fig10a.eps} \put(-3,50) {$(a)$} \put(30,16) {\textcolor{mygreen}{\footnotesize $\mathbf{x}_\sigma$}} \put(57,34) {\textcolor{blue} {\footnotesize $\mathbf{x}_l$}} \end{overpic} } % \vspace{0.4 cm} % \centerline{ \begin{overpic}[height=5cm, tics=10]{fig10b.eps} \put(-4,73) {$(b)$} \put(30,57) {\textcolor{mygreen}{\footnotesize $\mathbf{x}_c = \mathbf{x}_\sigma$}} \put(57,30) {\textcolor{blue} {\footnotesize $\mathbf{x}_c = \mathbf{x}_l$}} \end{overpic} \hspace{0.7 cm} \begin{overpic}[height=5cm, tics=10]{fig10c.eps} \put(-5,73) {$(c)$} \put(89.5,66) {\textcolor{black} {U}} \put(90,59) {\textcolor{black} {S}} \put(15,57) {\textcolor{blue} {\footnotesize $\mathbf{x}_c = \mathbf{x}_l$}} \put(45,32) {\textcolor{mygreen}{\footnotesize $\mathbf{x}_c = \mathbf{x}_\sigma$}} \end{overpic} } \caption{ $(a$) Locus of $\mathbf{x}_l$ (where $-\delta\rl/\rl$ is maximal) and $\mathbf{x}_\sigma$ (where $-\delta\sigma_{1,r}/|\sigma_{1,r}|$ is maximal) for a small control cylinder ($d=0.10D$) at $\Rey=40, 50 \ldots 120$. Solid lines show stabilising regions $\delta \sigma_{1,r} / \sigma_{1,r} \leq -1/2$ for two symmetric control cylinders at $\Rey=50, 60 \ldots 120$. $(b)$-$(c)$ Effect of two symmetric control cylinders ($d=0.10D$) located at $\mathbf{x}_l$ or at $\mathbf{x}_\sigma$: sensitivity analysis prediction for the normalized variation of $(a)$~recirculation length, $(b)$~growth rate of mode 1. The unshaded area corresponds to restabilisation of mode 1. } \label{fig:lose-correlation} \end{figure} All stabilising contours for $\Rey \geq 50$ are gathered in figure~\ref{fig:lose-correlation}$(a)$. The characteristic shrinking of the stabilising region is confirmed. Also shown for $\Rey \geq 40$ are the locus of two particular points: $\mathbf{x}_l$, where a control cylinder yields the largest $\rl$ reduction, and $\mathbf{x}_\sigma$, where a control cylinder has the maximal stabilising effect on mode 1. It can clearly be observed that $\mathbf{x}_l$ and $\mathbf{x}_\sigma$ move in opposite directions, with $\mathbf{x}_l$ eventually going outside the stabilising region. Quantitative values of $\delta \rl$ and $\delta \sigma_{1,r}$ are given in figures~\ref{fig:lose-correlation}$(b)$-$(c)$. With control cylinders located at $\mathbf{x}_l$, a recirculation length reduction of more than 35\% can be achieved for any $\Rey \leq 120$. (With $\mathbf{x}_c = \mathbf{x}_\sigma$, the reduction is of course not as large, and the recirculation length actually increases when $\Rey \gtrsim 65$.) As stressed previously, however, the corresponding controlled steady-state base flow will be observed only if stable, i.e. if $\delta \sigma_{1,r}/\sigma_{1,r} \leq 1/2$ (region labelled S). This is the case if $\mathbf{x}_c = \mathbf{x}_\sigma$ (at least for $\Rey \leq 120$). But when control cylinders are located at $\mathbf{x}_l$, although they do have a stabilising effect ($\delta \sigma_{1,r} \leq 0$), the latter is too small to restabilise the flow when $\Rey \gtrsim 70$. The same phenomenon is observed with sensitivity to wall forcing. Figure~\ref{fig:D_Uw_evgrowth-lc-ScalProd}$(a)$ shows the sensitivity of the most unstable growth rate to wall forcing. It compares very well with the results at $\Rey=60$ of \citet[figure 3 therein]{mar10-adjoint}. Unlike the sensitivity of recirculation length (figure \ref{fig:sensit_l_Uw-ntxy}) which keeps more or less the same structure at all Reynolds numbers, $\bnabla_{\mathbf{U}_w} \sigma_{1,r}$ varies substantially. In particular, at the top and bottom sides of the cylinder, wall-normal suction changes from largely stabilising to slightly destabilising. This translates into $\bnabla_{\mathbf{U}_w} \rl$ and $\bnabla_{\mathbf{U}_w} \sigma_{1,r}$ being very similar at $\Rey=40$ but very different at $\Rey=120$. Accordingly, the pointwise inner product of these two fields shown in figure~\ref{fig:D_Uw_evgrowth-lc-ScalProd}$(b)$ is positive everywhere on the cylinder wall at $\Rey=40$ but negative everywhere at $\Rey=120$. As a result, control configurations which shorten the recirculation length necessarily have a stabilising effect on mode 1 at low Reynolds numbers and a destabilising effect at higher Reynolds numbers. It should be noted that at $\Rey=100$, it is still possible to reduce $\rl$ and at the same time have a stabilising effect on mode 1 by using wall suction in a narrow region near $\theta=\pm\pi/2$. \begin{figure} \def\thisfigytop{25} \def\thisfigybot{-3} \psfrag{Scal Prod}[][][1][0]{$\bnabla_{\mathbf{U}_w} \rl \bcdot \bnabla_{\mathbf{U}_w} \sigma_{1,r}$} \psfrag{theta}[t][][1][0]{$\theta$} \psfrag{-pi} [][][1][0]{$-\pi$} \psfrag{-pi/2}[][][1][0]{$-\pi/2$} \psfrag{00} [][][1][0]{$0$} \psfrag{pi} [][][1][0]{$\pi$} \psfrag{pi/2} [][][1][0]{$\pi/2$} \vspace{0.9 cm} % \centerline{ \hspace{1 cm} \begin{overpic}[width=10.5cm,tics=10]{fig11a.eps} \put(-5, 25) {\footnotesize $(a)$} \put( 5, \thisfigytop) {\footnotesize $0.281$} \put(25.75, \thisfigytop) {\footnotesize $0.186$} \put(46.50, \thisfigytop) {\footnotesize $0.147$} \put(67.25, \thisfigytop) {\footnotesize $0.117$} \put(88, \thisfigytop) {\footnotesize $0.091$} % \put(2, \thisfigybot) {\footnotesize $\Rey=40$} \put(22.75, \thisfigybot) {\footnotesize $\Rey=60$} \put(43.50, \thisfigybot) {\footnotesize $\Rey=80$} \put(64.25, \thisfigybot) {\footnotesize $\Rey=100$} \put(85, \thisfigybot) {\footnotesize $\Rey=120$} \end{overpic} } % \vspace{0.9 cm} \centerline{ \begin{overpic}[width=10cm,tics=10]{fig11b.eps} \put(-2,61) {\footnotesize $(b)$} \put(48,45.5) {\footnotesize $\Rey=40$} \put(48,13) {\footnotesize $\Rey=120$} \put(26,54) {\footnotesize $\theta$} \put(22,50) {\footnotesize $x$} \put(13,60) {\footnotesize $y$} \end{overpic} } \caption{ $(a)$~Sensitivity of the leading growth rate to wall actuation $\bnabla_{\mathbf{U}_w} \sigma_{1,r}$. Flow is from left to right. Numbers correspond to the $L^\infty$ norm of $\bnabla_{\mathbf{U}_w} \sigma_{1,r}/\sigma_{1,r}$. % $(b)$~Pointwise inner product of $\bnabla_{\mathbf{U}_w} \rl$ (figure \ref{fig:sensit_l_Uw-ntxy}) and $\bnabla_{\mathbf{U}_w} \sigma_{1,r}$ along the cylinder wall. } \label{fig:D_Uw_evgrowth-lc-ScalProd} \end{figure} \section{Conclusions} \label{sec:conclusion} In this study, the sensitivity of recirculation length to steady forcing was derived analytically using a variational technique. Linear sensitivity analysis was applied to the two-dimensional steady flow past a circular cylinder for both subcritical and supercritical Reynolds numbers $40 \leq \Rey \leq 120$. Regions of largest sensitivity were identified: $\rl$ increases the most when small control cylinders are located close to the top and bottom sides of the main cylinder, and decreases the most when control cylinders are located farther downstream, outside the shear layers; regarding wall forcing, $\rl$ is most sensitive to normal blowing/suction at the sides of the cylinder. Validation against full non-linear Navier-Stokes calculations showed excellent agreement for small-amplitude control. Using linear stability analysis, it was observed that control configurations which reduce $\rl$ also have a stabilising effect on the most unstable eigenmode close to $\Rey_c$, both for bulk forcing and wall forcing. This property gradually disappears as $\Rey$ is increased. This is explained by the spatial structures of the sensitivities of $\rl$ and $\sigma_{1,r}$, which are very similar at low $\Rey$, but increasingly decorrelated at larger $\Rey$. Therefore, reducing the base flow recirculation length is an appropriate control strategy to restabilise the flow at moderate Reynolds numbers. At larger $\Rey$, aiming for an increase of $\rl$ is actually more efficient. In any case, one should keep in mind that results concerning $\rl$ reduction and obtained from a sensitivity analysis performed on the steady-state base flow are relevant only when the controlled flow is stable. To better control the flow in the supercritical regime, the sensitivity of the mean flow recirculation length should be considered. Not only would a method allowing to control the mean $\rl$ be interesting in itself, but targeting this important parameter of the mean state \citep{zie97,thi07} could also help stabilise the flow. This work is in progress, but the difficulty of such an approach lies in that the mean flow and the fluctuations are non-linearly coupled, which prevents the derivation of a simple expression for the sensitivity of the mean recirculation length. It would help, though, to determine whether the mean flow recirculation length in separated flows should be reduced or increased in order to mitigate the instability. This extension of the sensitivity analysis to the mean recirculation length could also include the effect of periodic excitation, a control strategy often used in turbulent flows \citep{Gre00,Gle02} and which can be interpreted as a mean-flow correction \citep{Sip12}. This study confirms the high versatility of Lagrangian-based variational techniques, which allow to compute the sensitivity of a great variety of quantities of interest in fluid flows. The recirculation length appears as a simple and relevant macroscopic parameter in separated flows, and can be targeted to design original control strategies. In the case of the cylinder flow, the fact that $\rl$ is a good proxy for flow stabilisation only up to a certain Reynolds number might be specific to the absolute nature of the instability in bluff body wakes \citep{mon88bluff}. These flows are typical examples of ``oscillators'', and are appropriately described by global linear stability analysis \citep{Cho05}. On the other hand, convectively unstable flows (or ``noise amplifiers''), such as separated boundary layers, usually exhibit large optimal transient growth (maximal energy amplification of an initial perturbation) and large optimal gain (maximal energy amplification from harmonic forcing to asymptotic response for a globally stable flow) as a result of the non-normality of the linearized Navier-Stokes operator \citep{Cho05}. One may wonder whether the recirculation length is more directly and strongly related to instability in such convectively unstable flows, and whether sensitivity analysis applied to $\rl$ would be efficient over a broader range of Reynolds numbers. The next step of this work is the application of a similar control strategy in wall-bounded separated boundary layer flows. \bigskip This work is supported by the Swiss National Science Foundation (grant no. 200021-130315). The authors are also thankful to the French National Research Agency (project no. ANR-09-SYSC-001). \bibliographystyle{jfm}
\section{Introduction} Many fundamental nuclear physics issues depend on our quantitative understanding of the $\beta$-decay phenomena in nuclei. Due to phase-space amplification effects, the $\beta$-decay rates are sensitive to both nuclear binding energies and $\beta$-strength functions. Within an appropriate $\beta$-decay model, the correct amount of the integral $\beta$-strength should be placed within the properly calculated $Q_{\beta}$- window provided that the spectral distribution is also close to the "true" $\beta$-strength function. It is desirable to have theoretical models which can describe the data wherever they can be measured, and predict the properties related to spin-isospin modes in the nuclei too short-lived to allow for experimental studies. One of the successful tools for studying charge-exchange nuclear modes is the quasiparticle random phase approximation (QRPA) with the self-consistent mean-field derived from a Skyrme-type energy-density functional (EDF), see e.g., \cite{inbsf96,ebnds99,inbsg00,bden02,colo07,bszzcx09,sag11,mb13}. These QRPA calculations enable one to describe the properties of the ground state and excited charge-exchange states using the same EDF. Experimental studies using the multipole decomposition analysis of the (n,p) and (p,n) reactions~\cite{w97,mda} found substantial Gamow-Teller (G-T) strength above the G-T resonance peak and have clarified a longstanding problem of the missing experimental G-T strength, hence resolving the discrepancies between the theoretical RPA predictions and the experimental measurements. It has been found necessary to take into account the coupling with more complex configurations in order to shift some strength to higher energies and to comply with the experimental results \cite{bh82,ks84,dosw87}. Using the Skyrme EDF and the RPA, such attempts in the past~\cite{cnbb94,csnbs98} have allowed one to understand the damping of charge-exchange resonances and their particle decay. Recently, the damping of the G-T mode has been investigated using the Skyrme-RPA plus particle-vibration coupling~\cite{ncbbm12}. The main difficulty is that the complexity of the calculations increases rapidly with the size of the configuration space and one has to work within limited spaces. Making use of the finite rank separable approximation (FRSA) \cite{gsv98,ssvg02,svg08} for the residual interaction one can perform Skyrme-QRPA calculations in very large two-quasiparticle spaces. Following the basic ideas of the quasiparticle-phonon model (QPM)~\cite{solo}, the approach has been generalized for the coupling between one- and two-phonon components of the wave functions~\cite{svg04}. The so-called FRSA was thus used to study the electric low-lying states and giant resonances within and beyond the QRPA ~\cite{svg08,svg04,sap12}. Recently, the FRSA approach was extended to charge-exchange nuclear excitations~\cite{svg12} and also for accommodating the tensor correlations which mimic the Skyrme-type tensor interactions~\cite{ss13}. In the present work we generalize the approach to the coupling between one- and two-phonon components in the wave functions. As an application of the method we study the $\beta$-decay half-lives of neutron-rich $N=50$ isotones and Ni isotopes and we compare to the most neutron-rich ($(N-Z)/A=0.28$) doubly-magic nucleus $^{78}$Ni which is also an important waiting point in the r-process~\cite{78ni}. In the case of $^{78}$Ni preliminary results of our calculation without the tensor interaction are already reported in Ref.~\cite{svbg13}. This paper is organized as follows. In Sec.~II, we sketch the method for including the effects of the phonon-phonon coupling. In Sec. III, we discuss the details of QRPA calculations for the $1^+$ states of the daughter nuclei and the $2^+$ states of the parent nuclei. In Sec.~IV, we analyze the results of the calculations of $\beta$-decay rates. Conclusions are finally drawn in Sec.~V. \section{The FRSA model} The FRSA model for charge-exchange excitations was already introduced in Refs.~\cite{svg12,ss13}. In the present study, this method is extended by including the coupling between one- and two-phonon terms in the wave functions of G-T states. The starting point is the Hartree-Fock(HF)-BCS calculation~\cite{RingSchuck} of the parent ground state within a spherical symmetry assumption. In the particle-hole (p-h) channel we use the Skyrme interaction with the triplet-even and triplet-odd tensor components introduced in Refs.~\cite{s59,sbf77}. The continuous part of the single-particle spectrum is discretized by diagonalizing the HF hamiltonian on a harmonic oscillator basis. The inclusion of the tensor interaction results in the following modification of the spin-orbit potential in coordinate space~\cite{hsplb07,t44}: \begin{equation} \label{sop} U^{(q)}_{S.O.}=\frac{W_0}{2r}\left(2\frac{d\rho_q}{dr}+\frac{d\rho_{q'}}{dr}\right) +\left(\alpha\frac{J_q}{r}+\beta\frac{J_{q'}}{r}\right), \end{equation} where $\rho_q$ and $J_q$ ($q=n,p$) are the densities and the spin-orbit densities, respectively. The coefficients $\alpha$ and $\beta$ can be separated into contributions from the central force ($\alpha_c$, $\beta_c$) and the tensor force ($\alpha_T$, $\beta_T$)~\cite{hsplb07,t44}. The pairing correlations are generated by the density-dependent zero-range force \begin{equation} \label{pair} V_{pair}({\bf r}_1,{\bf r}_2)=V_{0}\left( 1-\eta\left(\frac{\rho \left( r_{1}\right) } {\rho _{0}}\right)^{\gamma}\right) \delta \left( {\bf r}_{1}-{\bf r}_{2}\right), \end{equation} where $\rho _{0}$ is the nuclear saturation density. The values of $V_{0}$, $\eta$ and $\gamma$ are fixed to reproduce the odd-even mass difference of the studied nuclei~\cite{svg08}. To build the QRPA equations on the basis of HF-BCS quasiparticle states with the residual interactions consistently derived from the Skyrme EDF in the p-h channel and from the zero-range pairing force in the particle-particle (p-p) channel is a standard procedure~\cite{t05}. The dimensions of the QRPA matrix grow very rapidly with the size of the nuclear system unless severe and damaging cut-offs are made to the 2-quasiparticle configuration space. It is well known that, if the QRPA matrix elements take a separable form the QRPA energies can be obtained as the roots of a relatively simple secular equation~\cite{solo,BB}. In the case of the Skyrme interaction this feature has been exploited by different authors \cite{gsv98,s99,n02}. The main step of the FRSA is to simplify the central p-h interaction $V^{C}_{ph}$ by approximating it by its Landau-Migdal form. All Landau parameters with $l > 1$ are equal to zero in the case of Skyrme interactions. We keep only the $l=0$ terms in $V^{C}_{ph}$ and this approximation is reasonable~\cite{gsv98,svg08}. The two-body Coulomb residual interaction is dropped. Therefore we can write $V^{C}_{ph}$ as \begin{eqnarray} V^{a}_{res}({\bf r}_1,{\bf r}_2)=N_0^{-1}\left[ F_0^{a}(r_1)+G_0^{a}(r_1) {\bf \sigma}_1\cdot{\bf \sigma}_2 \right.\nonumber\\\left.+(F_0^{'a}(r_1)+G_0^{'a}(r_1){\bf \sigma }_1\cdot{\bf \sigma}_2){\bf \tau }_1\cdot{\bf \tau }_2\right] \delta ({\bf r}_1-{\bf r }_2), \label{res.int} \end{eqnarray} where $a$ is the channel index $a=\{ph,pp\}$; ${\bf \sigma}_i$ and ${\bf \tau}_i$ are the spin and isospin operators, $N_0 = 2k_Fm^{*}/\pi^2\hbar^2$ with the Fermi momentum $k_F$ and $m^{*}$ is the nucleon effective mass. The coefficients $G^{pp}_0$ and $G^{'pp}_0$ are zero, while the expressions for $F^{ph}_0$, $F^{'ph}_0$, $G^{ph}_0$, $G^{'ph}_0$ and $F^{pp}_0$, $F^{'pp}_0$ can be found in Ref.~\cite{sg81} and in Ref.~\cite{svg08}, respectively. For the case of electric excitations one can neglect the spin-spin terms since they play a minor role. Though it is well known that the p-p interaction in the spin-isospin channel ($T=0$ pairing) suppresses the $\beta^{-}$-decay half-lives, in the present study we assume $G^{'pp}_0$=0 in order to separate the sole impact of the tensor force. As proposed in Refs.~\cite{bzzxsc09,ss13}, we simplify the tensor p-h interaction by replacing it by the two-term separable interaction, where the strength parameters are adjusted to reproduce the centroid energies of the G-T and spin-quadrupole strength distributions calculated with the original tensor p-h interaction. The p-h matrix elements and the antisymmetrized p-p matrix elements can be written in a separable form in the space of the angular coordinates~\cite{gsv98,svg08}. After integrating over the angular variables the one-dimensional radial integrals are numerically calculated by choosing a large enough cutoff radius $R$ and using an $N$-point integration Gauss formula with abscissas ${r_k}$ and weights ${w_k}$~\cite{gsv98}. Thus, one is led to deal with a problem where the matrix elements of the residual interaction are sums of products and the number $\tilde N$ of terms in the sums depends only on $N$. In particular, $\tilde N=4N+4$ and $\tilde N=6N$ are obtained for the cases of G-T and electric excitations, respectively. One can call it a separable approximation of finite rank $\tilde N$ since finding the roots of the secular equation amounts to find the zeros of a $\tilde N \times \tilde N$ determinant, and the dimensions of the determinant are independent of the size of the configuration space, i.e., of the nucleus considered. The studies of Refs. \cite{svg08,svg12} enable us to conclude that $N$=45 is enough for the electric and charge-exchange excitations considered here in nuclei with $A\le 208$. In the next step, we construct the wave functions from a linear combination of one-phonon and two-phonon configurations \begin{eqnarray} \Psi _\nu (J M) = \left(\sum_iR_i(J \nu )Q_{J M i}^{+}\right. \nonumber\\ \left.+\sum_{\lambda _1i_1\lambda _2i_2}P_{\lambda _2i_2}^{\lambda _1i_1}( J \nu )\left[ Q_{\lambda _1\mu _1i_1}^{+}\bar{Q}_{\lambda _2\mu _2i_2}^{+}\right] _{J M }\right)|0\rangle~, \label{wf} \end{eqnarray} where $Q_{\lambda\mu i}^{+} |0\rangle$ ($\bar{Q}_{\lambda\mu i}^{+} |0\rangle$) is the G-T (electric) excitation having energy $\omega_{\lambda i}$ ($\bar{\omega}_{\lambda i}$). The normalization condition for the wave functions~(\ref{wf}) is \begin{equation} \sum\limits_iR_i^2( J \nu)+ \sum_{\lambda _1i_1 \lambda _2i_2} (P_{\lambda _2i_2}^{\lambda _1i_1}(J \nu))^2=1. \end{equation} The amplitudes $R_i(J \nu)$ and $P_{\lambda_2i_2}^{\lambda_1i_1}(J \nu)$ are determined from the variational principle which leads to a set of linear equations \begin{eqnarray} (\omega_{\lambda i}-\Omega_\nu )R_i(J \nu ) +\sum_{\lambda _1i_1 \lambda_2i_2} U_{\lambda _2i_2}^{\lambda _1i_1}(J i) P_{\lambda_2i_2}^{\lambda _1i_1}(J \nu )=0, \label{2pheq1} \end{eqnarray} \begin{eqnarray} (\omega _{\lambda _1i_1}+\bar{\omega}_{\lambda _2i_2}-\Omega_\nu )P_{\lambda _2i_2}^{\lambda _1i_1}(J \nu)\nonumber\\ +\sum\limits_i U_{\lambda _2i_2}^{\lambda _1i_1}(J i)R_i(J \nu )=0. \label{2pheq2} \end{eqnarray} The rank of the set of linear equations (\ref{2pheq1}) and (\ref{2pheq2}) is equal to the number of one- and two-phonon configurations included in the wave function (\ref{wf}). For its solution it is required to compute the Hamiltonian matrix elements coupling one- and two-phonon configurations \begin{equation} U_{\lambda _2i_2}^{\lambda _1i_1}(J i)= \langle 0| Q_{J i } H \left[ Q_{\lambda _1i_1}^{+}\bar{Q}_{\lambda _2i_2}^{+}\right] _{J} |0 \rangle. \end{equation} Eqs.~(\ref{2pheq1}) and (\ref{2pheq2}) have the same form as the QPM equations~\cite{ks84,solo}, where the single-particle spectrum and the residual interaction are derived from the same Skyrme EDF. In the allowed G-T approximation, the $\beta^{-}$-decay rate is expressed by summing the probabilities of the energetically allowed G-T transitions (in units of $G_{A}^{2}/4\pi$) weighted with the integrated Fermi function \begin{eqnarray} T_{1/2}^{-1}=\sum_m \lambda^{m}_{if} = D^{-1}\left(\frac{G_{A}}{G_{V}}\right)^{2}\times\nonumber\\ \sum_{m}f_{0}(Z,A,E_i-E_{1_m^+})B(G-T)_{m}, \label{t} \end{eqnarray} \begin{equation} E_i-E_{1_m^+}\approx\Delta M_{n-H}+\mu_n-\mu_p-E_m, \label{et} \end{equation} where $\lambda^{m}_{if}$ is the partial $\beta^{-}$-decay rate, $G_A/G_V$=1.25~\cite{Suhonen} is the ratio of the weak axial-vector and vector coupling constants and $D$=6147 s (see Ref.~\cite{Suhonen}). $\Delta M_{n-H}=0.782$~MeV is the mass difference between the neutron and the hydrogen atom, $\mu_n$ and $\mu_p$ are the neutron and proton chemical potentials, respectively, $E_i$ is the ground state energy of the parent nucleus, and $E_{1_m^+}$ denotes a state of the daughter nucleus $(Z,A)$. $E_m$ and $B(G-T)_{m}$ are the solutions either of the QRPA equations, or of Eqs.~(\ref{2pheq1})-(\ref{2pheq2}) taking into account the two-phonon configurations. Thus, to calculate the half-lives by Eqs.(\ref{t}) and (\ref{et}) an approximation worked out in Ref.~\cite{ebnds99} is used. It allows one to avoid an implicit calculation of the nuclear masses and $Q_{\beta}$-values. However, one should realize that the related uncertainty in constraining the parent nucleus ground state calculated with the chosen Skyrme interaction is transferred to the values of the neutron and proton chemical potentials entering Eq.~(\ref{et}). \section{Details of calculations} \begin{figure}[t!] \includegraphics[width=1.0\columnwidth]{be2n50} \caption{Energies and $B(E2)$ values for up-transitions to the $[2_1^+]_{QRPA}$ states in the neutron-rich $N=50$ isotones. Results of the calculations without the tensor interaction (open triangles) and with the tensor interaction (filled triangles) are shown. Experimental data (filled diamonds) are taken from Ref.~\cite{pbsh13}.} \end{figure} \begin{figure}[t!] \includegraphics[width=1.0\columnwidth]{be2ni} \caption{Same as Fig. 1, but for the neutron-rich Ni isotopes} \end{figure} We apply the approach to study the influence of the coupling between one- and two-phonon terms in the wave functions, as well as the tensor force effects on the strength distributions of G-T states in the neutron-rich Ni isotopes and $N=50$ isotones. To obtain the interaction in the p-h channel, we use the Skyrme interaction SGII~\cite{sg81} and the zero-range tensor interaction of Ref.~\cite{bzzxsc09} with $\alpha_T$=-180~MeVfm$^5$ and $\beta_T$=120~MeVfm$^5$ for the parameters $\alpha$ and $\beta$ of the one-body spin-orbit potential of Eq.(\ref{sop}). Since the SGII parametrization gives reasonable values for the Landau parameters $F^{'}_{0}=0.73$ and $G^{'}_{0}=0.50$, one obtains a successful description of the spin-dependent properties, in particular for the experimental energies of the G-T resonances of $^{90}$Zr~\cite{sg81}. In Ref.~\cite{ss13}, the FRSA model has been used to calculate the G-T states of $^{90}$Zr and $^{208}$Pb, and the FRSA results reproduce the main features of the G-T strength distributions obtained with the exact treatment of Skyrme tensor interactions in the RPA. For the interaction in the p-p channel, we use a zero-range volume force, i.e., $\eta=$0 in Eq.~(\ref{pair}) and $V_0$=-270 MeVfm$^{3}$ with a smooth cut-off at 10 MeV above the Fermi energies~\cite{svg08,k90}. This value of the pairing strength has been fitted to reproduce the experimental pairing energies of $^{70,72,74,76}$Ni obtained from binding energies of neighbouring nuclei. This choice of the pairing interaction has also been used for a satisfactory description of the experimental data of $^{90,92}$Zr and $^{92,94}$Mo~\cite{sap12}. Because of the closed $Z=28$ and $N=50$ shells, the $T=0$ pairing is not effective in these nuclei~\cite{nmvpr05} and, therefore, it can be neglected. We now use the FRSA of the residual interaction and carry out QRPA calculations in very large two-quasiparticle spaces. In particular, the cut-off in the discretized continuous part of the s.p. spectra is at 100~MeV. Because of the inclusion of the tensor correlation effects within the $1p-1h$ and $2p-2h$ configuration space, we do not need any quenching factor~\cite{bh82}. For the ansatz of the wave function (\ref{wf}), the Ikeda sum rule $S_{-}-S_{+}=3(N-Z)$ is fulfilled, see e.g., \cite{ks84}. Our configurational space is sufficient to exhaust this sum rule for the G-T strength of the nuclei that we studied without and with the tensor force. As an illustrative example, in the case taking into account the tensor correlations for $^{82}_{32}$Ge we obtain $S_{-}=56.85$ and $S_{+}=2.84$ in the QRPA, the inclusion of the 2p-2h effects gives the same values. The results without the tensor interaction indicate $S_{-}=54.46$ and $S_{+}=0.45$. Let us now focus on the properties of the low-energy G-T state, since we are interested in the $\beta$-decay half-lives. Its experimentally known half-life puts an indirect constrain on the calculated G-T strength distributions within the $Q_{\beta}$-window. Let us examine the extension of the configuration space to one- and two-phonon terms. To construct the wave functions~(\ref{wf}) of the low-lying $1^{+}$ states we use only the $[1^{+}_{i}\otimes\lambda^{+}_{i'}]_{QRPA}$ terms and all electric phonons with $\lambda > 2$ vanish, i.e., G-T phonons from the charge-exchange modes are only used in the two-phonon terms, as in Ref.~\cite{cnbb94,ncbbm12}. All one- and two-phonon configurations with the transition energies $|E_{1_m^+}-E_i|$ up to 10~MeV are included. We have checked that the inclusion of the high-energy configurations leads to minor effects on the half-life values. It is interesting to study the energies, reduced transition probabilities and the structure of the $2_1^+$ QRPA state. The calculated $2_1^+$ energies and transition probabilities in the neutron-rich $N=50$ isotones are compared with existing experimental data~\cite{pbsh13} in Fig.1. The FRSA model with the tensor interaction reproduces the experimental data~\cite{pbsh13} very well. We find that the tensor interaction induces a reduction of the $2_1^+$ collectivity and it results in a decrease of the transition probability, see Fig.1. There is a remarkable increase of the $2_1^+$ energy of $^{78}$Ni in comparison with those in $^{76}$Fe and $^{80}$Zn. It corresponds to a standard evolution of the $2_1^+$ energy near closed shells. As can be seen from Fig.2, the behavior of $2_1^+$ energies of $^{76,78,80}$Ni is similar to that of the $N=50$ isotones. It is seen that the inclusion of the tensor interaction does not change energies and transition probabilities along this Ni isotopic chain except for $^{70}$Ni. Including the tensor interaction changes contributions of the main configurations only slightly, but the general structure of the $2_1^+$ state remains the same. The neutron amplitudes are dominant in all Ni isotopes and the contribution of the main neutron configuration $\{1g_{9/2},1g_{9/2}\}$ decreases from 89\% for $^{70}$Ni to 77\% for $^{76}$Ni when neutrons fill the subshell $1g_{9/2}$. Thus, a satisfactory description of $2_1^+$ energies is found for $^{72,74,76}$Ni. Our calculated $2_1^+$ energy of $^{70}$Ni is about a factor of 2 too large compared to the data value~\cite{pbsh13}. This is likely due to the overestimation of the neutron contribution to the $2_1^+$ wave function within the QRPA. One can expect some improvement if the coupling with the two-phonon components of the wave functions is taken into account~\cite{svg04,sap12}. \begin{figure}[t!] \includegraphics[width=1.0\columnwidth]{lambda80zn} \caption{The phonon-phonon coupling effect on the $\beta$-transition rates in $^{80}$Zn. The left and right panels correspond to the calculations within the QRPA and taking into account the $2p-2h$ configurations, respectively. Results of the calculations without (resp. with) the tensor interaction are shown in the upper (resp. lower) panels.} \end{figure} \begin{figure}[t!] \includegraphics[width=1.0\columnwidth]{lambda72ni} \caption{Same as Fig. 3, but for $^{72}$Ni.} \end{figure} \begin{figure}[t!] \includegraphics[width=1.0\columnwidth]{halfliven50} \caption{The phonon-phonon coupling effect on $\beta^-$-decay half-lives of the neutron-rich $N=50$ isotones. Results of the calculations without the tensor interaction (open triangles, circles, squares) and with the tensor interaction (filled triangles, circles, squares) are shown. The squares correspond to the half-lives calculated with inclusion of the phonon-phonon coupling, the triangles are the QRPA calculations. Results including the $[1^{+}_{i}\otimes 2^{+}_{i'}]_{QRPA}$ configurations are denoted by the circles. Experimental data (filled diamonds) are from Refs.~\cite{expthalflives,Xu}.} \end{figure} \begin{figure}[t!] \includegraphics[width=1.0\columnwidth]{halfliveni} \caption{Same as Fig. 5, but for the neutron-rich Ni isotopes. Experimental data are taken from Refs.~\cite{expthalflives,Xu,6874ni}.} \end{figure} \section{Results for $\beta$-decay rates} Using the same set of parameters we calculate the low-lying G-T strength distributions of the neutron-rich nuclei $^{76}_{26}$Fe, $^{70,72,74,76,78,80}_{\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,28}$Ni, $^{80}_{30}$Zn, $^{82}_{32}$Ge and $^{84}_{34}$Se. First, the properties of the low-lying $1^+$ states of the daughter nuclei are studied within the one-phonon approximation. As expected, the largest contribution ($>$88\%) in the calculated $\beta^-$-decay half-life comes from the $[1_1^+]_{QRPA}$ state. To illustrate it, the $\beta$-transition rates $\lambda^{m}_{if}$ of $^{80}$Zn and $^{72}$Ni are shown in Fig.~3 and Fig.~4, respectively. The transition energies $E_{1^+_m}-E_i$ refer to the ground state of the parent nucleus. QRPA results calculated without the tensor interaction indicate that the dominant configuration of the $[1_1^+]_{QRPA}$ states is $\{\pi2p\frac{3}{2}\nu2p\frac{1}{2}\}$ whose contribution is about 91~$\%$ in all the cases considered. In other words, the $[1_1^+]_{QRPA}$ state is not collective and, therefore, the $\beta^-$-decay is related to the unperturbed $\{\pi2p\frac{3}{2}\nu2p\frac{1}{2}\}$ energy. At the same time the calculated $\log ft$ value increases from 3.7 for $^{72}$Ni to 3.9 for $^{80}$Ni. The $\beta^-$-decay half-lives of the neutron-rich $N=50$ isotones and the Ni isotopes are shown in Fig.~5 and Fig.~6, respectively. The calculated values overestimate the experimental data~\cite{78ni,6874ni,expthalflives,Xu}. Moreover, for $^{70}$Ni, $^{82}$Ge and $^{84}$Se, the $[1_1^+]_{QRPA}$ state of the daughter nucleus is above the parent ground state, i.e., this calculation predicts stable $^{70}$Ni, $^{82}$Ge and $^{84}$Se. It is worth pointing out that the relativistic QRPA approach~\cite{nmvpr05} gives a similar structure of the $[1_1^+]_{QRPA}$ state for these nuclei. As a general trend, we observe a redistribution of the G-T strengths by the inclusion of the tensor interaction in the QRPA, in the same way as it was found in Refs.~\cite{bszzcx09,sag11,ss13}. The tensor correlations shift up about 10\% of the G-T strength to the energy region above 30MeV. Also, the tensor interaction makes a downward shift of the strength in the G-T resonance region below the peak energy and the $[1_1^+]_{QRPA}$ state is moved downwards. This is illustrated in the cases of $^{80}$Zn and $^{72}$Ni (see left bottom panels of Figs.~3-4). For $^{80}$Ni, the $[1_1^+]_{QRPA}$ energy shift reaches 2.7~MeV. In particular, because of this shift, we get unstable $^{70}$Ni and $^{82}$Ge. In addition, the tensor correlations lead to a collective structure for the $[1_1^+]_{QRPA}$ state with the dominance of the $\{\pi2p\frac{3}{2}\nu2p\frac{1}{2}\}$ configuration. In the case of the $N=50$ isotones the $\{\pi2p\frac{3}{2}\nu2p\frac{1}{2}\}$ contribution increases from 46\% in $^{78}$Ni to 72\% in $^{84}$Se, whereas for Ni isotopes this contribution decreases from 57\% in $^{70}$Ni to 39\% in $^{80}$Ni. The $[1_1^+]_{QRPA}$ collectivity is reflected in the $\log ft$ value and there is a slight decrease of the $\log ft$ from 2.3 for $^{70}$Ni to 2.0 for $^{80}$Ni. As can be seen from Fig.5 and Fig.6, the half-lives calculated with the tensor force are about 60-2000 times shorter than those calculated without. This analysis within the one-phonon approximation can help to identify the tensor correlation effects, but it is only a rough estimate. It is worth to mention that the first discussion of the strong impact of the tensor correlations on the $\beta^-$-decay half-lives based on QRPA calculations with Skyrme forces has been done in Ref.~\cite{mb13}. Let us now discuss the extension of the space to one- and two-phonon configurations in the FRSA model when the tensor interaction is taken into account. In all ten nuclei, the dominant contribution in the wave function of the first $1^+$ state comes from the $[1^+_1]_{QRPA}$ configuration, but the two-phonon contributions are appreciable. The main two-phonon components of the $1^+_1$ wave function are the $[1_1^+\otimes2_1^+]_{QRPA}$ and $[1_1^+\otimes0_2^+]_{QRPA}$ configurations. When the tensor interaction is not included, the main configuration is only the $[1_1^+\otimes2_1^+]_{QRPA}$ configuration. As a result, the inclusion of the two-phonon terms results in an increase of the transition energies $|E_{1_1^+}-E_i|$ and the energy shift is large (1.1 MeV for $^{82}$Ge and 0.8 MeV for $^{80}$Zn) in comparison with 0.2 MeV in the case of the doubly-magic nucleus $^{78}$Ni. The $\log ft$ value is practically unchanged. One can see from Figs.5-6 that the effects of the phonon-phonon coupling produce a sizable impact on the $\beta$-decay half-life which is reduced by a factor 30. Specifically, the $[1_1^+\otimes2_1^+]_{QRPA}$ configuration is the important ingredient for the half-life description since the $[2_1^+]_{QRPA}$ state is the lowest collective-electromagnetic excitation which leads to the minimal two-phonon energy and the maximal matrix elements coupling one- and two-phonon configurations of Eqs.~(\ref{2pheq1}) and (\ref{2pheq2}). Figs. 5 and 6 also show the half-life reduction as an effect of the quadrupole-phonon coupling. For $^{84}$Se, the experimental half-life is $T_{1/2}=3.26\pm0.10$~min\cite{expthalflives}, but our model predicts that this nucleus is stable against $\beta$-decay and, in particular, $E_{1_1^+}-E_i$=0.0~MeV. One can expect an improvement if the $T=0$ pairing interaction is taken into account. We intend to extend our formalism to include the $T=0$ pairing effects. It is interesting to discuss the change of the half-lives along the chains of the $N=50$ isotones and the Ni isotopes. As pointed out in Refs.~\cite{expthalflives,6874ni,Xu}, there are different evolutions of the existing experimental half-lives, namely, the 37.3-time reduction of half-life values from $^{82}$Ge to $^{78}$Ni and the gradual reduction of half-lives with increasing neutron number for the Ni isotopes. One can see that our results reproduce this behaviour which is sensitive to the isotopic and isotonic dependences of the transition energies $|E_{1_1^+}-E_i|$. We find that the energy $|E_{1_1^+}-E_i|$ for $^{82}$Ge ($^{74}$Ni) is 3.1 (1.2) times less than that in $^{78}$Ni. It is worth to mention that the calculated half-life of $^{82}$Ge is in reasonable agreement with the experimental data. As can be seen from Fig. 6, the main discrepancies between measured and calculated half-lives of Ni isotopes are due to too strong tensor correlations and one should seek for improvements in the tensor part of the effective interaction used. Finally, for neighbours of $^{78}$Ni we give predictions as the bottom limit of the half-life (0.7 ms for $^{76}$Fe and 2.5 ms for $^{80}$Ni). It is seen that the phonon-phonon coupling plays a minor role. Our calculated half-life of $^{80}$Ni is in qualitative agreement with the recently observed value of $23.9^{+26.0}_{-17.2}$ms~\cite{Xu}. The SGII parametrization of the central force gives a reasonable description of properties of the G-T and charge-exchange spin-dipole resonances of $^{90}$Zr~\cite{sag11,sg81} and the low-energy spectrum of the quadrupole excitations for nuclei near $^{90}$Zr~\cite{sap12}. For the half-life description, the quantitative agreement with the experimental data is not satisfactory for the neutron-rich nuclei. A possible reason might be the underestimated symmetry energy of 26.8~MeV in the case of the SGII set. A half-life study of the influence of the two-phonon terms with taking into account the tensor interaction for the different parametrizations of central Skyrme force is still underway. Inclusion of the tensor interaction results in significant increase of the transition energies and partial rates of the main G-T transitions. Including the $2p-2h$ configurations leads to further increase of its rate and to the appearance of the weak fragmented satellites at low transition energy (see, e.g., right bottom panels of Figs.3-4). Importantly, such a change of the G-T strength distribution takes place in the near-threshold region. Thus, an additional constraint on the $\beta$-strength distribution is also given by delayed neutron emission probability ($P_n$-value)~\cite{pn05}. Since the FRSA model enables one to evaluate the coupling of QRPA phonons to more complex configurations, such calculations that take into account the $2p-2h$ fragmentation of the QRPA excitations are now in progress. \section{Conclusions} Starting from a Skyrme effective interaction the G-T strength in the $Q_{\beta}$- window has been studied within the extended FRSA model including both the tensor interaction and $2p-2h$ configurations effects. The suggested approach enables one to perform the calculations in very large configuration spaces. Using the parameter set SGII+tensor interaction, we have applied this model to the G-T states in the neutron-rich Ni isotopes and $N=50$ isotones, for which experimental data of $\beta^{-}$-decay rates are available. The inclusion of the tensor interaction leads to a redistribution of the G-T strength. The low-energy G-T strength is fragmented and the $1^+_1$ state is moved downwards. We observe that the coupling between one- and two-phonon terms with the $2^+$ phonon states is strong in these nuclei, and the $2p-2h$ fragmentation and damping of the QRPA excitations are thus important. In particular, the energy shift of the lowest $1^+$ state due to the phonon-phonon coupling is large: 0.8 MeV in $^{80}$Zn and 1.1 MeV in $^{82}$Ge in comparison with 0.2 MeV in the case of $^{78}$Ni. Taking into account these effects results in a dramatic reduction of $\beta$-decay half-lives. It is shown that the $2p-2h$ impact on the half-lives comes inherently from the $[1_1^+\otimes2_1^+]_{QRPA}$ term of the wave function of the $1^+_1$ state. At a qualitative level, our results reproduce the experimental evolution of the half-lives, i.e., for the $N=50$ isotones we describe the sharp reduction of half-lives with decreasing proton number and at the same time, for the Ni isotopes, the gradual reduction of half-lives with increasing neutron number. We give predictions for open-shell nuclei $^{76}$Fe and $^{80}$Ni near $^{78}$Ni that are important for stellar nucleosynthesis. Using the strong tensor correlations~\cite{bzzxsc09} our estimation is rather the bottom limit of these half-lives. Also, our calculations show that the influence of the phonon coupling is small on the half-lives of $^{76}$Fe and $^{80}$Ni. More information on the $\beta$-strength distributions can be derived from simultaneous analysis of the half-lives and the $P_n$ values~\cite{pn05,Miernik}. Such a program of detailed calculations is underway, the results will be reported in connection to the RIKEN experiments on $^{78}$Ni~\cite{RIKEN}. \section*{Acknowledgments} A.P.S., V.V.V., I.N.B and N.N.A thank the hospitality of IPN-Orsay where the part of this work was done. This work is partly supported by the IN2P3-RFBR agreement No.~110291054 and the IN2P3-JINR agreement.
\section{Context} The issue of whether the baryonic material in galaxy disks does, or does not, contribute most of the central attraction is important for many reasons, such as understanding the dynamical structure of galaxies and comparison with predictions from galaxy formation models. While the total central attraction is determined by the rotation curve, the relative contributions of baryonic and dark matter are not easily separated. Various indirect arguments have been advanced to bear on this issue, one of which is that strong bars in sub-maximal disks should be slowed by dynamical friction \citep{DS00}. Since $R_c$, the radius of corotation for a bar, increases as the bar slows, we proposed an observationally accessible measure of whether a bar has been slowed by friction as the value of the dimensionless ratio ${\cal R} = R_c/a_B$, where $a_B$ is the semi-major axis of the bar. We argued that a strong bar can remain fast, ${\cal R} \la 1.4$, only if the barred disk is close to maximal. However, this argument was called into question in a recent paper by \citet{Atha14} who reported $\cal R$ values from a suite of 15 simulations with initial disk gas fractions ranging from 0 to 100\% in halos that were axisymmetric, as well as similar models with mildly or strongly triaxial halos. She concluded ``The models, by construction, have roughly the same azimuthally averaged circular velocity curve and halo density and they are all submaximal, {i.e.}\ according to previous works they are expected to have all roughly the same ${\cal R}$ value, well outside the fast bar range ($1.2 \pm 0.2$). Contrary to these expectations, however, these simulations end up having widely different ${\cal R}$ values, either within the fast bar range, or well outside it. This shows that the ${\cal R}$ value can not constrain the halo density, nor determine whether galactic discs are maximal or submaximal.'' Here we show that these statements reflect a simple misunderstanding of the criterion, and that her results appear to be in good agreement with it. \begin{figure*} \begin{center} \includegraphics[width=.27\hsize, angle=270]{fast_slow_3.ps} \medskip \includegraphics[width=.25\hsize, angle=270]{gas_mass_R1_h1_h2_h3.eps} \end{center} \caption{Above, Figure 1 reproduced from \citet{Atha14}, showing the time evolution of her measurements of ${\cal R}$ in all 15 of her simulations. Below, Figure 16 from \citet{AMR13} showing the time evolution of the gas+new star particles inside $R=1\;$kpc in the same simulations. The red and magenta lines relate to the same models in both rows of figures, but the blue and green lines relate to models that have been interchanged between the two rows.} \end{figure*} \bigskip\bigskip\bigskip \section{The evidence} The claims by \citet{Atha14} were based on calculations that were described more fully in \citet{AMR13}. They all began with identically the same mass distribution, except that the disk was composed of different fractions of gas and star particles, and the spherical halo in one set was distorted into mildly and more strongly triaxial shape in the other two sets. At the outset, the disk was clearly submaximal, since the peak circular speed from the disk was fractionally less than the circular speed from the halo at the same radius. The simulations lacking any gas, formed large ($a_B \sim 10\;$kpc), strong bars that were fiercely braked by dynamical friction against the halo, so that ${\cal R} \gg 1.4$ by the end of the simulation. It is clear from Fig.~1 of \citet{Atha14}, reproduced here in the upper panel of our Fig.~1, that the final ${\cal R}$ values decreased systematically with increasing gas fraction. However, as shown in the lower panel of our Fig.~1, which is reproduced from Fig.~16 of \citet{AMR13}, the mass distribution in the disks with gas was rearranged within the first 2~Gyr of evolution, well before the bars had formed and settled. The mass that accumulated in the inner disk increased with the gas mass fraction, and in the cases that began with 100\% gas (magenta lines) the mass within 1~kpc increased some 4-fold over that at the start. The other simulations start with smaller gas fractions, and larger stellar fractions (the initial stellar fraction is omitted in their Figure), and therefore the relative increase in the total inner disk mass is proportionately less. This initial rearrangement of the mass distribution is sufficient to change the dynamical properties of the models. For example, if the 100\% gas disk were to have contracted homologously to an exponential disk of half the radial scale but the same total mass, then the central surface density would increase 4-fold, have a peak in the circular speed that is $\surd2$ times higher than before lying at a radius that is half that of the original peak. From Fig.~1 of AMR13, that would mean the disk contribution reaches $\sim200\;$km/s at $R=3.3\;$kpc, which would clearly be maximal. The contraction is unlikely to be homologous, but this crude approximation is perhaps consistent with the initially gas-rich models forming shorter bars. The size of the bar that forms in a simulation is determined by a variety of factors, including the steepness of the inner rise in the rotation curve and the distribution of mass in the disk. The bar in the model with 100\% gas has $a_B \sim 5\;$kpc in the spherical halo \citep[][her Fig.~2]{Atha14} about half that in the stars only case, and the decreasing bar sizes can also be seen in the snaphots of Figs.~4 \& 5 and the measurements in Fig.~9 of \citet{AMR13}. It is therefore no surprise that the shorter bars in the increasingly dominant disks experience weaker friction, such that ${\cal R}$ manifests the trend shown in the upper panel of Fig.~1. It seems very likely that the bars that experience little frictional drag are effectively in maximum disks. Note also that the initial rise in the central gas density was greater in the models with triaxial halos. The principal consequence of triaxiality is to promote the inflow of gas, which is borne out by the mass increase by $t=2\;$Gyr in all cases except for the 100\% gas case, which is less in the strongly triaxial case than in the mildly. This odd result could simply be stochasticity \citep{SD06} that could be checked in multiple runs with different random seeds. \section{Conclusions} Once the mass rearrangement, due to the initial shrinking size of the gas component, is taken into account, the simulations reported by \citet{AMR13} and \citet{Atha14} appear to be consistent with all previous work \citep[see][for a review]{Sell14}. The different $\cal R$ values at the ends of her 15 simulations can be understood as reflecting the differences in the effective degree of disk domination in the inner parts of the models. Thus it would appear that her results in agreement with the conclusion that only maximal disk models can have ${\cal R} \sim 1.2 \pm 0.2$ after some period of evolution. The statements to the contrary in \citet{Atha14} seem to reflect a simple misunderstanding by that author of what exactly is the criterion. In fact, it relates to the mass distribution in the model at the time of the measurement, which is the only observationally accessible measurement, of course. Her mis-statements derive from an incorrect expectation that models that begin from the same mass distribution, but with differing fractions of gas, should all experience the same dynamical friction; in fact, the mass rearrangement in the early part of the evolution makes the gas-rich disks more nearly maximal, leading to correspondingly weaker friction on the bar that subsequently forms.
\section{Introduction} In the last fifty years, the physics of disordered systems has turned out to be tremendously rich, and the field is still offering challenging and unexpected results. Among those, the manifestations of weak and strong (Anderson) localization of coherent waves are paradigmatic examples \cite{Anderson58, Bergmann84}. In practice, Anderson localization very often manifests itself as a halt of wave transport. This signature has been largely exploited in a number of experiments searching for Anderson localization of classical waves in disordered media \cite{Schwartz07, Hu08, Maret12} or of matter waves subjected to time-periodic \cite{Chabe08, Lemarie10} and random \cite{Billy08, Roati08, Kondov11, Jendrzejewski12, Semeghini14} optical potentials. At the same time, recent works \cite{Cherroret12, Karpiuk12, Micklitz14, Loon14} pointed out that in ultracold-atom setups, the momentum distribution of a matter wave in a random potential can exhibit a highly nontrivial dynamics due to localization if it is initially launched with a nonzero mean wave-vector $\bs{k}_0$. The scenario is then the following. First, over a time scale of the order of the Boltzmann transport mean free time $\tau_B$, an isotropization of the distribution takes place as particles' momenta are being randomized by the disorder \cite{Plisson13}. During this process, a narrow coherent backscattering (CBS) peak emerges around the direction $-\bs{k}_0$. After a few $\tau_B$, this peak gains a maximum visibility and sits on top of a broader isotropic, ring-shaped distribution \cite{Cherroret12}. At a later time, a second interference peak appears in the forward direction $+\bs{k}_0$ \cite{Karpiuk12}. The visibility of this coherent forward scattering (CFS) peak increases slowly, and finally reaches a maximum value at a time of the order of the Heisenberg time $\tau_H$, defined as the inverse of the mean spacing between the energy levels of the system (see below for a more precise discussion). Beyond $\tau_H$, the system no longer evolves and the asymptotic distribution has the central symmetry (if time-reversal invariance is preserved), with two identical CBS and CFS peaks. Experimentally, the CBS effect of ultracold atoms has been recently observed \cite{Josse12}, motivating further investigations like the sensitivity of CBS to external dephasing \cite{Cord14}. An observation of the CFS effect is, on the other hand, still missing. From a theoretical point of view, while the physics of CBS is today well understood -- it stems from wave amplitudes travelling along the same multiple scattering sequence but in opposite directions \cite{Aegerter09} -- the mechanism of CFS is much less obvious. In fact, the building of the CFS peak relies on interference sequences where particles are scattered back and forth several times between the directions $-\bs{k}_0$ and $+\bs{k}_0$ \cite{Karpiuk12, Micklitz14, Loon14}. While this mechanism is inefficient in a purely diffusive system, it becomes strongly enhanced when the wave gets confined in a limited region of space: wave interference can then accumulate and make the CFS peak macroscopic. The situation typically occurs in an infinite system if Anderson localization comes into play (this was the scenario originally considered in \cite{Karpiuk12}), but also, as discussed in the present paper, when the wave is trapped within a limited region of space of size smaller than the localization length. In this respect, a full characterization of the CFS effect is required in order to unambiguously attribute its appearance to Anderson localization. In this paper, we present a detailed study of the momentum distribution of a matter wave launched in a random potential, bringing special attention to the CFS effect to which we propose a systematic numerical analysis combined with theoretical predictions. We consider a random potential of speckle type, routinely used in current experiments with ultracold atoms \cite{ColdDisorderRevs}, and focus on the two-dimensional (2D) geometry for which the CFS peak clearly distinguishes itself from the isotropic part of the momentum distribution \cite{Karpiuk12} (we refer the reader to \cite{Loon14, Micklitz14} for the one-dimensional and quasi one-dimensional cases). The main concepts discussed in the paper as well as the theoretical framework are introduced in Sec. \ref{framework}. In Sec. \ref{sectionL}, we analyze the building up of the CFS peak for a matter wave scattered diffusively in a limited volume of size much smaller than the localization length of the problem, $L\ll\xi$. Then, in Sec. \ref{sectionxi}, we address the more difficult but more interesting case $\xi\ll L$ where the CFS peak is triggered by the confinement of the matter wave stemming from Anderson localization. This configuration is typically the one of experiments, where an atomic wave packet is initially prepared and evolves in the presence of the disorder, without any confining box. This experimental scenario, studied in Sec.~\ref{feasibility}, has however two additional ingredients: (i) the matter wave has a broad energy distribution for which both localized and diffusive atoms coexist and (ii) the initial state has a finite spatial width, affecting the shape and height of the CFS peak. Despite these complications, we show that, in the absence of an artificial confining box, the diffusive components studied in Sec.~\ref{sectionL} do not contribute to the CFS signal, and that the latter can be observed and used as a ``smoking gun'' of Anderson localization. In Sec. \ref{Conclusion}, we finally summarize our results and discuss some open questions. \section{Momentum distribution of a matter wave in a 2D random potential} \label{framework} \subsection{Numerical experiment} \label{numerics} In order to introduce the physics discussed in the paper, let us first consider a simple numerical experiment. We start from a plane matter wave $|\bs{k}_0\rangle$ and propagate it with the evolution operator $\exp(-i\hat{H}t)$, where $\hat{H}=\bs{p}^2/(2m)+V(\textbf{r})$ with $V(\textbf{r})$ a 2D random potential (from here on we set $\hbar=1$). Following recent experiments on ultracold atoms, we choose $V(\textbf{r})$ to be a blue-detuned speckle potential with mean value $\av{V(\textbf{r})}=0$ and correlation function $\av{V(\textbf{r})V(\textbf{r}')}=[2V_0J_1(|\textbf{r}-\textbf{r}'|/\zeta)/(|\textbf{r}-\textbf{r}'|/\zeta)]^2$, with $\zeta$ the correlation length. This potential is numerically generated in a standard way, by convoluting a circular Gaussian random field with a cutoff function that simulates the diffusive plate used in experiments \cite{Huntley89, Horak98}. As soon as the random potential is turned on, $\bs{k}_0$ is no longer a good quantum number and the system starts to evolve. The time propagation is achieved on a 2D grid of size $L\times L$ with periodic boundary conditions along $x$ and $y$, by using an iterative method based on the expansion of the evolution operator in combinations of Chebyshev polynomials of the Hamiltonian \cite{Fehske09, Roche97}. In the simulations, a cell of surface $(\pi\zeta)^2$ is discretized in typically 8-10 steps along both $x$ and $y$. After the evolution, the momentum distribution is calculated by applying a discrete Fourier transformation on the final wave function. This procedure is repeated for many disorder realizations, which finally gives access to the disorder-averaged momentum distribution. Throughout the paper, lengths, momenta, energies and times will be given in units of $\zeta$, $\zeta^{-1}$, $1/(m\zeta^2)$ and $m\zeta^2$, respectively. A typical distribution obtained at long times (here $t=10^3$) is shown in Fig. \ref{twinpeaks}, for $V_0=5$, $k_0\equiv|\bs{k}_0|=1.5$ and for a system size $L=20\pi$. \begin{figure} \includegraphics[width=0.72\linewidth]{2D_mon.eps} \caption{(Color online) Density plot of the long-time momentum distribution obtained after numerical propagation of a plane wave $|\bs{k}_0\rangle$ in a 2D speckle potential. Parameters are $k_0=1.5$, $V_0=5$, $L=20\pi$ and $t=10^3$, where lengths, momenta, energies and times are given in units of $\zeta$, $\zeta^{-1}$, $1/(m\zeta^2)$ and $m\zeta^2$, respectively. The left peak is due to coherent backscattering and the right peak to coherent forward scattering. Data are averaged over 7200 disorder realizations. } \label{twinpeaks} \end{figure} The distribution of Fig. \ref{twinpeaks}, $\overline{n}(\bs{k})=\overline{n}_\text{D}+\overline{n}_\text{CBS}+\overline{n}_\text{CFS}$, exhibits three components: an isotropic, diffusive ring-shaped ``background'' $\overline{n}_\text{D}$ and two interference peaks $\overline{n}_\text{CFS}$ and $\overline{n}_\text{CBS}$ centered at $\pm \bs{k}_0$. The ring describes the quasi-elastic isotropization of atomic momenta in the course of the propagation in the random potential. The peak centered at $-\bs{k}_0$ is the coherent backscattering peak, and the peak centered at $+\bs{k}_0$ is the coherent forward scattering peak \cite{Cherroret12, Karpiuk12}. The stationary distribution of Fig. \ref{twinpeaks} is obtained after a long time of propagation in the random potential. Before this final state establishes however, the system explores a number of regimes summarized in Fig. \ref{phase_diagram}, and that we discuss below. \subsection{Diagrammatic theory in the diffusive regime} \label{theory_diag} At short times, a diagrammatic description of the momentum distribution can be developed. We briefly summarize below the essential steps of this approach. In momentum space, the wave function $|\psi(t)\rangle$ at time $t$ reads \begin{equation} \langle\bs{k}|\psi(t)\rangle=\int \frac{d^2\bs{k}' dE_1}{(2\pi)^3} e^{-iE_1 t} \langle\bs{k}|\hat{G}^R(E_1)|\bs{k}'\rangle \langle\bs{k}'|\phi\rangle, \end{equation} where $|\phi\rangle$ is the wave function at time $t=0$ and $\hat{G}^R(E_1)=(E_1-\hat{H}+i0^+)^{-1}$ is the retarded Green's operator at energy $E_1$. The disorder-averaged momentum distribution at time $t$, $\av{n}(\bs{k},t)=\av{|\langle\bs{k}|\psi(t)\rangle|^2}$, then involves the disorder-averaged intensity propagation kernel $\av{\langle\bs{k} |\hat{G}^R(E_1)|\bs{k}'\rangle\langle\bs{k}''| \hat{G}^A(E_2)|\bs{k}'\rangle}$, integrated over the momenta $\bs{k}'$ and $\bs{k}''$ and over the energies $E_1$ and $E_2$ [$\hat{G}^A(E_1)=(E_1-\hat{H}-i0^+)^{-1}$ is the advanced Green's operator]. Since the disorder average restores translation invariance and thus momentum conservation, this kernel takes the simpler form $(2\pi)^2\delta(\bs{k}'-\bs{k}'')\Phi_{\bs{k}'\bs{k} E}(\omega)$, where $E=(E_1+E_2)/2$, $\omega=E_1-E_2$, and where $\Phi_{\bs{k}'\bs{k} E}$ remains to be determined. The disorder-averaged momentum distribution at time $t$ thus reads \cite{Kuhn07} \begin{equation} \av{n}(\bs{k},t) = \intdone{\omega}e^{-i\omega t} \int \frac{d^2\bs{k}' dE}{(2\pi)^3} \Phi_{\bs{k}'\bs{k} E}(\omega)n_0(\bs{k}'), \end{equation} where $n_0(\bs{k}')=| \langle\bs{k}'|\phi\rangle|^2$. In the numerical simulation, we consider for simplicity an initial plane wave $|\phi\rangle = |\bs{k}_0\rangle$, leading to $n_0(\bs{k}')=(2\pi)^2\delta(\bs{k}'-\bs{k}_0)$ and to \begin{equation} \label{rhokprimet} \av{n}(\bs{k},t) = \intdone{\omega} e^{-i\omega t}\intdone{E}\Phi_{\bs{k}_0\bs{k} E}(\omega). \end{equation} In this section and the next one, we discuss the dynamics at a given energy $E$. To lighten the notations we therefore drop the $E-$dependence of all time and length scales, keeping in mind that the full momentum distribution is a superposition of many energy components [see Eq. (\ref{rhokprimet})], each of which behaving \emph{a priori} differently in the disorder. We also assume that disorder is weak, namely $k_0\ell_B\gg 1$, where $\ell_B$ is the Boltzmann transport mean free path \cite{AM}. In this limit, the various time scales of the system are well separated, and several regimes of transport can be clearly identified, as we now discuss. \begin{figure}[h] \includegraphics[width=0.75\linewidth]{diagrams.eps} \caption{Leading-order diagrams contributing to the momentum distribution. (a): series of ladder diagrams giving rise to the isotropic part of the distribution, Eq. (\ref{Background}). (b): series of crossed diagrams giving rise to the CBS peak, Eq. (\ref{CBS_diag}). The two series of diagrams (c) and (d) give the main contribution to the CFS peak in the diffusive regime $\tau_B\ll t\ll \text{min}(L,\xi)^2/D_B$, Eq. (\ref{CFS_diag}). They equally contribute for a time-reversal invariant system. } \label{diagrams} \end{figure} \begin{figure} \includegraphics[width=0.75\linewidth]{phase_diagram.eps} \caption{(Color online) Schematic dynamical phase diagram of a 2D, weakly disordered system. $\tau_B$ is the transport mean free time and $\ell_B$ the transport mean free path. For $t\lesssim \tau_B$ atoms undergo a few scattering events and the distribution gets isotropized (regime 1). For $\tau_B\ll t\ll \tau_D, \tau_\text{loc}$, the randomization process is completed and transport is diffusive (regime 2). Beyond the Thouless time $\tau_D\equiv L^2/D_B$ (for $L<\xi$) or the localization time $\tau_\text{loc}\equiv \xi^2/D_B$ (for $L>\xi$), a particle passes through regions already visited (``ergodic limit'', regime 3). Finally, for times much larger than the Heisenberg time $\tau_H\equiv2\pi\overline{\nu} \text{min}(L,\xi)^2$ the particle has resolved the discreteness of energy levels and the system no longer evolves (``quantum limit'', regime 4). } \label{phase_diagram} \end{figure} At lowest order in $(k_0\ell_B)^{-1}\ll1$, only pairs of trajectories following exactly the same multiple scattering sequence, i.e. with no net phase difference, survive the disorder average. The corresponding contribution to the kernel $\Phi_{\bs{k}_0\bs{k} E}(\omega)$, the so-called series of ladder diagrams, is shown in Fig. \ref{diagrams}(a). It describes a diffusion mechanism and leads to a fast isotropization process of atomic momenta at very short times $t\sim \tau_B$ where $\tau_B$ is the Boltzmann transport time. This is the regime 1 in Fig. \ref{phase_diagram} and it has been analyzed in \cite{Plisson13} by means of a kinetic approach. In the following we will not consider this regime, focusing only on times $t\gg\tau_B$ where the isotropization process is completed and diffusion is fully established (regime 2 in Fig. \ref{phase_diagram}). For $t\gg\tau_B$, the $\omega-$dependence of the diagram in Fig. \ref{diagrams}(a) is purely controlled by its central (ladder) part, given by the diffusion propagator $P_E(\bs{q},\omega)=1/(-i\omega+D_B\bs{q}^2)$ (with $D_B$ the Boltzmann diffusion constant) at zero momentum \cite{Cherroret12}: \begin{eqnarray} \Phi_{\bs{k}_0\bs{k} E}(\omega)&=& 2 \av{\langle\bs{k}_0|\text{Im}\hat{G}^R(E)|\bs{k}_0\rangle}\times 2\av{\langle\bs{k}|\text{Im}\hat{G}^R(E)|\bs{k}\rangle} \nonumber\\ &&\times P_E(0,\omega)/[2\pi\overline{\nu}(E)], \end{eqnarray} where the two average Green's operators come from the `legs' of the diagram and where $\overline{\nu}(E)$ is the average density of states per unit volume at energy $E$. Inserting this expression into Eq. (\ref{rhokprimet}) and performing the Fourier integral over $\omega$, we obtain the time-independent, isotropic and ring-shaped contribution to the disorder-averaged momentum distribution, well visible in Fig. \ref{twinpeaks}: \begin{equation}\label{Background} \av{n}_\text{D}(\textbf{k}) =\av{n}_\text{D}(k)=\intdone{E}\dfrac{A(\bs{k},E)A(\bs{k}_0,E)}{2\pi\overline{\nu}(E)}, \end{equation} where we have introduced the spectral function $A(\bs{k},E)=2\pi\av{\langle\bs{k}|\delta(E-\hat{H})|\bs{k}\rangle}=-2\av{\langle\bs{k}|\text{Im}\hat{G}^R(E)|\bs{k}\rangle}$ \cite{Shapiro12}. Note that $\av{n}_\text{D}(\textbf{k})$ is indeed isotropic because the spectral function only depends on $k\equiv |\bs{k}|$ \cite{Shapiro12}. The physical interpretation of Eq.~(\ref{Background}) is rather clear: the spectral function $A(\bs{k}_0,E)$ describes the probability density that the initial state with momentum $\bs{k}_0$ has an energy $E$, while $A(\bs{k},E)/[2\pi\overline{\nu}(E)]$ in turn describes the probability density that a state with energy $E$ has a momentum $\bs{k}.$ Note that one could imagine a slightly different situation where the initial state is not a plane wave, for exemple a wave packet with finite size, see Sec.~\ref{feasibility}, or a more complicated state obtained after the disordered potential is progressively switched on. The analysis developed in this paper can be used to describe such a variant, simply by replacing $A(\bs{k}_0,E)$ by the energy distribution of the initial state. The series of crossed diagrams (b) in Fig. \ref{diagrams} gives a correction to Eq. (\ref{Background}) that describes the CBS peak growing around the backscattering direction $\bs{k}=-\bs{k}_0$. Its calculation follows the same lines as that of diagram (a) \cite{Cherroret12}. Exactly at backscattering $\bs{k}=-\bs{k}_0$, the CBS contribution reaches rapidly a stationary value given by \begin{equation}\label{CBS_diag} \av{n}_\text{CBS}(-\bs{k}_0)=\intdone{E}\dfrac{A^2(\bs{k}_0,E)}{2\pi\overline{\nu}(E)}. \end{equation} According to Eqs. (\ref{Background}) and (\ref{CBS_diag}), at backscattering the momentum distribution is exactly twice the value of the isotropic background, $\overline{n}(-\bs{k}_0)=\overline{n}_\text{D}(k_0)+\overline{n}_\text{CBS}(-\bs{k}_0)=2\overline{n}_\text{D}(k_0)$, which is an emblematic signature of the CBS effect resulting from a plane-wave source \cite{Aegerter09}. Finally, it was shown in \cite{Karpiuk12, Micklitz14} that at short enough times, the leading contribution to the CFS peak is given by the diagrams (c) and (d) in Fig. \ref{diagrams}, which combine two successive crossed and ladder sequences. These diagrams are peaked in the forward direction, unlike other diagrams of the same order of magnitude in perturbation theory, which provide flat contributions to the momentum distribution \cite{Karpiuk12}. Diagram (d) is obtained from diagram (c) by time-reversing one of the complex amplitudes. Since the system we consider has the time-reversal symmetry, both diagrams equally contribute, giving \begin{eqnarray}\label{CFS_diag} \av{n}_\text{CFS}(\bs{k}_0)&\simeq&2\intdone{E}\intdone{\omega} e^{-i\omega t}\dfrac{A^2(\bs{k}_0,E)}{[2\pi\overline{\nu}(E)]^2\tau_s}\times\nonumber\\ &&\intdtwo{\bs{q}}A(\bs{q},E)P_E(\bs{q}+\bs{k}_0,\omega)^2. \end{eqnarray} In two dimensions, Eq. (\ref{CFS_diag}) yields $\av{n}_\text{CFS}(\bs{k}_0,t)/\av{n}_\text{D}(k_0)\sim 1/(k_0\ell_B)$, which is a constant, small contribution for weak disorder \cite{Karpiuk12} (note that this result is different from the case of a one-dimensional or quasi one-dimensional geometry, for which the corresponding Eq. (\ref{CFS_diag}) gives $\av{n}_\text{CFS}(\bs{k}_0,t)\sim\sqrt{t}$ \cite{Micklitz14}). This means that the CFS peak is hardly visible in the diffusive regime. As shown in \cite{Karpiuk12}, Eq. (\ref{CFS_diag}) is only valid at short enough times, when higher-order corrections to the CFS peak remain small. The question of the description of CFS at longer times, where such corrections cannot be neglected anymore, is the object of the next section.\\ \subsection{Theoretical description of the momentum distribution in the ergodic and quantum regimes} For a 2D disordered system of size $L$ and characterized by a localization length $\xi$, Eqs. (\ref{Background}), (\ref{CBS_diag}) and (\ref{CFS_diag}) strictly speaking hold only in the diffusive regime where $\tau_B\ll t\ll \text{min}(L,\xi)^2/D_B$ (regime $2$ in Fig. \ref{phase_diagram}). When $L<\xi$, $L^2/D_B\equiv\tau_D$ is the so-called Thouless time, i.e. the typical time needed by an atom to reach the boundary of the system \cite{Thouless74}. In the opposite limit $\xi< L$, $\xi^2/D_B\equiv \tau_\text{loc}$ can be interpreted as the localization time, i.e. the time scale at which atoms become sensitive to Anderson localization. In both cases, as soon as $t>\tau_D$ or $\tau_\text{loc}$, atoms start to feel that they are confined in a finite region of space: this is the ergodic regime, see Fig. \ref{phase_diagram}. In the ergodic regime the system still evolves, until it eventually resolves its spectrum (``quantum limit'', regime 4 in Fig. \ref{phase_diagram}). This happens at a time scale known as the Heisenberg time, $\tau_H\equiv2\pi\overline{\nu}\text{min}(L,\xi)^2$, which is the inverse of the mean level spacing in a volume of size $L^2$ (when $L<\xi$) or $\xi^2$ (when $L>\xi$). Note that in two dimensions the ratios $\tau_H/\tau_D,\ \tau_H/\tau_\text{loc}\sim k_0\ell_B$ are very large in the weak-disorder limit assumed here, such that the ergodic regime is typically very broad. In the ergodic regime, atoms pass again and again through spatial regions they have already explored. This phenomenon produces a a highly non-perturbative accumulation of interference, and many corrections to the CFS diagrams in Fig. \ref{diagrams}(c) and \ref{diagrams}(d) become relevant and come into play. At first sight, one might think of summing up all these corrections by directly ``chaining'' an arbitrary number of series of crossed and ladder diagrams. This approach seems however hopeless because of the rapid proliferation of the number of such corrections at longer and longer times (the Hikami boxes connecting the series can be dressed in many possible ways \cite{Karpiuk12}). Nevertheless, a description of the transport dynamics in the ergodic regime can be obtained from the supersymmetric nonlinear $\sigma$-model developed by Efetov \cite{Efetov99}. Within this approach, the intensity propagator has the general form \begin{widetext} \begin{equation} \label{Susy} \Phi_{\bs{k}_0\bs{k} E}(\omega)=\dfrac{A(\bs{k}_0,E)A(\bs{k},E)}{4L^2}\int DQ\left[Q_{15}(0)Q_{51}(0)+ Q_{11}(\bs{k}-\bs{k}_0)Q_{55}(\bs{k}_0-\bs{k})+ Q_{35}(\bs{k}+\bs{k}_0)Q_{53}(-\bs{k}-\bs{k}_0)\right]e^{-F[Q]}. \end{equation} \end{widetext} In Eq. (\ref{Susy}), $\int (...)DQ$ is a functional integral over a $8\times 8$ supermatrix $Q$ that fulfills the constraint $Q^2=1$, and $\Lambda=\text{diag}(\mathbb{1}_4,-\mathbb{1}_4)$. The elements of $Q$ are complex and Grassmann fields (the Hamiltonian here belongs to the orthogonal symmetry class). The action of the $\sigma$-model is $F[Q]=[\pi\overline{\nu}(E)/4]\text{Str}\int d^2\textbf{r}[-D(\boldsymbol\nabla Q)^2-2i\omega\Lambda Q]$, where Str denotes the supertrace. The elements of $Q$ are arranged in two retarded and advanced sectors describing the product of the two Green's functions involved in $\Phi_{\bs{k}_0\bs{k} E}$. These sectors are split in two sectors containing variables and their complex conjugate (pertaining to the time-reversal symmetry), themselves being split in two bosonic and fermionic sectors required to perform the disorder average \cite{Efetov99, Mirlin00}. At this stage, Eq. (\ref{Susy}) is general, with the only restriction that time should be larger than $\tau_B$ and disorder should be weak. After integration over $Q$ the three terms in the right-hand-side give rise to functions of $\bs{k}$ which are respectively (from left to right) constant, peaked around $\bs{k}=\bs{k}_0$, and peaked around $\bs{k}=-\bs{k}_0$, and can thus \emph{a posteriori} be identified as the isotropic background of the distribution, the CFS and the CBS peaks. Due to the complicated manifold spanned by the matrix $Q$ however \cite{Efetov99, Mirlin00}, these functions can only be calculated in a few specific cases. In the next section, we focus on the limit $L\ll\xi$ where an exact result can be obtained for the CFS contrast in the ergodic regime $t>\tau_D$, and even in the long-time, quantum regime $t\gg\tau_H$, where the momentum distribution is expected to reach its final, stationary form. Finally, it is worthwhile to note that whatever the ratio $L/\xi$ but at times $\tau_B\ll t\ll \text{min}(L,\xi)^2/D_B$ (diffusive regime), the field integrals over $Q$ can also be performed, using a perturbation theory around the high-frequency saddle point $Q=\Lambda$ \cite{Micklitz14}. The calculation of the three terms in Eq. (\ref{Susy}) in this limit then reproduces Eq. (\ref{Background}) for the background, Eq. (\ref{CBS_diag}) for the CBS peak at $\bs{k}=-\bs{k}_0$, and Eq. (\ref{CFS_diag}) for the CFS peak at $\bs{k}=\bs{k}_0$. This in particular confirms the conjecture that at short times the series of diagrams in Figs. \ref{diagrams}(c) and \ref{diagrams}(d) are indeed those responsible for the CFS effect \cite{Karpiuk12}. \section{CFS in a limited volume ($L\ll\xi$)} \label{sectionL} \subsection{CFS in the ergodic and quantum regimes} In this section, we assume $L\ll \xi$ and focus on times $t\gg \tau_D$ (regions $2$ and $3$ in Fig. \ref{phase_diagram}). Since $L\ll\xi$, the confinement effect leading to the CFS peak stems from the finite volume of the system. Its dynamics can be accessed from Eq. (\ref{Susy}) by replacing the functional integral by a definite one and using the parametrization of the matrix $Q$ proposed by Efetov \cite{Efetov99} (``zero-dimensional'' approximation). With this strategy, the CFS contribution to Eq. (\ref{Susy}) yields, at $\bs{k}=+\bs{k}_0$, \begin{equation} \label{CFS_longt} \av{n}_\text{CFS}(\bs{k}_0,t)=\intdone{E}\dfrac{A^2(\bs{k}_0,E)}{2\pi\overline{\nu}(E)}2\pi\overline{\nu}(E)L^2K_E(t), \end{equation} where \begin{eqnarray} \label{KE_GOE} &&2\pi\overline{\nu}(E)L^2K_E(t)=\\ && \left \{\begin{array}{l} (t/\tau_H)[2-\ln(1+2t/\tau_H)],\ \tau_D\ll t\leq\tau_H \\ 2-(t/\tau_H)\ln[(1+2t/\tau_H)/(2t/\tau_H-1)],\ t\geq\tau_H, \nonumber\\ \end{array} \right. \end{eqnarray} with $\tau_H\equiv2\pi\overline{\nu}(E)L^2$ the Heisenberg time associated with the system of size $L$. This result shows that starting from $t=\tau_D$, the CFS peak slowly increases until a few Heisenberg times as atoms keep exploring the volume of the system. For $t\gg \tau_H$, $2\pi\overline{\nu}(E)K_E(t)\simeq 1$ and the CFS peak has reached its maximum. In this limit, the momentum distribution no longer evolves in time because atoms have resolved the discreteness of energy levels. It is interesting to note that the function $K_E(t)$, as given by Eq. (\ref{KE_GOE}), is nothing but the so-called form factor -- the Fourier transform of the correlation of density-of-states fluctuations -- of the Gaussian Orthogonal Ensemble of random matrices \cite{Mirlin00} (see also Sec. \ref{sectionxi}). This shows in particular that the CFS peak is intrinsically connected with the spectral properties of the disordered system \cite{Loon14}. In real space, the form factor also governs the dynamics of the ``mesoscopic echo effect'', i.e. the enhancement of the probability for a spatially narrow wave packet to return to the origin in the presence of disorder \cite{Prigodin94}. Finally, we mention for completeness that the CBS peak at $\bs{k}=-\bs{k}_0$ as well as the isotropic component of the distribution [the third and first terms in Eq. (\ref{Susy}), respectively] can also be derived from Eq. (\ref{Susy}) for $t>\tau_D$, using the zero-dimensional approximation. This calculation eventually leads to the same expressions given in Eqs. (\ref{Background}) and (\ref{CBS_diag}), signaling that these formulas in fact hold very generally, not only in the diffusive regime but also in the long-time limit $t>\tau_D$ \subsection{Numerical simulations} \label{numerics_WD} For $L\ll\xi$, we now have a complete physical picture of the dynamics, with Eqs. (\ref{Background}), (\ref{CBS_diag}) and (\ref{CFS_longt}) describing respectively the isotropic background, the CBS and the CFS peaks. In order to test the validity of these formulas, we perform extensive numerical simulations of the time-resolved momentum distribution of a matter wave in a 2D speckle potential, using the approach outlined in Sec. \ref{numerics}. To achieve the condition $L\ll\xi$, we set $k_0=2$ and consider a relatively weak value of the disorder amplitude, $V_0=1$. For these parameters, we compute numerically the spectral function $A(\bs{k}_0,E)=-2\text{Im}\overline{\langle\bs{k}_0|\hat{G}(E)|\bs{k}_0\rangle}$, where $\langle\bs{k}_0|\hat{G}(E)|\bs{k}_0\rangle=-i\int_0^\infty dt\langle\bs{k}_0|\text{exp}[i(E-\hat{H})t] |\bs{k}_0\rangle$ is obtained by propagation of the plane-wave state $|\bs{k}_0\rangle$. The spectral function is shown in the main panel of Fig. \ref{SpectralF}, giving an estimation of the energy distribution of the matter wave. Its shape is reminiscent of the Lorentzian expected in the limit of weak disorder \cite{AM}. The inset of Fig. \ref{SpectralF} also shows the energy dependence of the localization length, which we compute numerically using the transfer-matrix technique in two dimensions \cite{footnote1, McKinnon83, Slevin14, Delande14}. At a given $L$ corresponds a certain energy $E_L$, below which atoms are typically localized [$\xi(E<E_L)<L$] and above which atoms are typically diffusive [$\xi(E>E_L)>L$]. For the largest value of $L$ considered in this section ($L=25\pi$, see below) we find $E_L\simeq0.05$. This value falls in the left tail of the spectral function, where the latter is almost zero. \begin{figure}[h] \includegraphics[width=1\linewidth]{SpectralF1.eps} \caption{ Spectral function $A(\bs{k}_0,E)$ as a function of energy, obtained from numerical simulations of plane-wave propagation in a 2D speckle potential, for $V_0=1$ and $k_0=2$ (lengths, momenta and energies are respectively given in units of $\zeta$, $\zeta^{-1}$ and $1/(m\zeta^2)$, where $\zeta$ is the correlation length of the random potential). The inset shows the localization length given by the transfer-matrix approach as a function of energy, for the same value of $V_0$. Data are averaged over $16$ disorder realizations. The energy $E_L$ corresponding to a localization length $\xi(E_L)=L$ for $L=25\pi$ is indicated: energies below $E_L$ are typically localized [$\xi(E)<L$], while energies above $E_L$ are typically diffusive [$\xi(E)>L$]. From the main plot, it is seen that below $E=E_L$ the spectral function is almost zero, which means that for these parameters essentially all atoms are diffusive.} \label{SpectralF} \end{figure} This means that for $V_0=1$ and $k_0=2$ essentially all atoms fulfill the condition $L<\xi$. We show in Fig. \ref{background} a radial cut along $k_x=0$ of the momentum distribution obtained numerically after a propagation time $t=7200$, i.e. well beyond the Boltzmann transport mean free time which is $\tau_B\simeq 7$ at $E=E_0\equiv k_0^2/(2m)$. This plot is expected to describe the radial shape of the isotropic part of the distribution, given by Eq. (\ref{Background}). In Fig. \ref{background} we also show this theoretical prediction, in which we used the numerically computed spectral functions $A(\bs{k}_0,E)$, $A(\bs{k},E)$, and density of states $\overline{\nu}(E)\equiv\int d\bs{k}/(2\pi)^3A(\bs{k},E)$. We see that the theory perfectly matches the numerical results without any adjustable parameter. \begin{figure} \includegraphics[width=0.9\linewidth]{background_V1.eps} \caption{(Color online) Cut along $k_x=0$ of the momentum distribution, describing the radial shape of its isotropic, ring-shaped part. Points were obtained from numerical simulations of plane-wave propagation in a 2D speckle potential in the limit $L\ll\xi$ ($V_0=1$ and $k_0=2$), after averaging over $6000$ disorder realizations. $L$ is set to $25\pi$. The red curve is the theoretical prediction (\ref{Background}), in which spectral functions and density of states are computed numerically. There is no adjustable parameter.} \label{background} \end{figure} Let us now focus on the contrast of the CBS and CFS peaks with respect to the isotropic background, defined respectively as $C_\text{CBS}\equiv \overline{n}_\text{CBS}(-\bs{k}_0,t)/\overline{n}_\text{D}(k_0,t)$ and $C_\text{CFS}\equiv\overline{n}_\text{CFS}(\bs{k}_0,t)/\overline{n}_\text{D}(k_0,t)$. According to Eqs. (\ref{Background}) and (\ref{CBS_diag}), we have evidently \begin{equation} \label{CCBS} C_\text{CBS}=1,\ t\gg\tau_B. \end{equation} Furthermore, by comparing Eqs. (\ref{Background}) and (\ref{CFS_longt}) and assuming that the energy dependence of $\tau_H$ is smooth as compared to that of the spectral function (which is a very good approximation for the relatively low value of $V_0$ considered in this section), we have \begin{equation} \label{CCFS} C_\text{CFS}\simeq2\pi\overline{\nu}(E_0)L^2K_{E_0}(t),\ t\gg\tau_D, \end{equation} where the expression of $K_E(t)$ is given by Eq. (\ref{KE_GOE}). Eqs. (\ref{CCBS}) and (\ref{CCFS}) are shown in the main panel of Fig. \ref{CBSCFS_weak} as a function of time (dashed and solid curves, respectively), together with the CBS and CFS contrasts obtained from our numerical simulations of plane-wave propagation (green and red symbols, respectively). Circles were obtained for a system size $L=15\pi$, squares for $L=20\pi$ and crosses for $L=25\pi$. The agreement between analytical formulas and the numerics is excellent. Note in particular that all the numerical points fall on the same master curve when plotted as a function of $t/\tau_H$, which confirms the universal scaling in $t/L^2$ of the function $K_{E_0}(t)$. The inset additionally shows the CBS and CFS contrasts together with the isotropic, background contribution to the momentum distribution at short times $t\ll\tau_H$: both the background and the CBS contrast become time independent after a few $\tau_B$, see Eqs. (\ref{Background}) and (\ref{CBS_diag}), while the CFS contrast increases slowly, linearly in time, in agreement with the small-time limit $2\pi\overline{\nu}(E_0)L^2K_{E_0}(t\ll \tau_H)\simeq 2t/\tau_H$. \begin{figure} \centering\includegraphics[width=1\linewidth]{CBSCFS_weak.eps} \caption{(Color online) Main panel: contrast of the CBS (upper green symbols) and CFS (lower red symbols) peaks as a function of $t/\tau_H$ [here $\tau_H=\tau_H(E_0)=2\pi\overline{\nu}(E_0) L^2$], obtained from numerical simulations of plane-wave propagation in the limit $L\ll\xi$ ($V_0=1$ and $k_0=2$). Circles were obtained for $L=15\pi$, squares for $L=20\pi$ and crosses for $L=25\pi$. Data are averaged over $6000$ disorder realizations and over a time window $\Delta t=250$. The dashed and solid curves are the theoretical predictions (\ref{CCBS}) and (\ref{CCFS}), respectively. Inset: contrasts at short times $t\ll\tau_H$. For comparison, the scaled background $\overline{n}_D(k_0,t)/\overline{n}_D(k_0,\infty)$ is also shown (blue crosses). Both the background and the CBS contrast become time independent after a few $\tau_B$, while the CFS contrast increases slowly, linearly in time. } \label{CBSCFS_weak} \end{figure} \section{CFS in a localized system ($\xi\ll L$)} \label{sectionxi} \subsection{CFS peak at long times} We now consider the case $\xi\ll L$, where the confinement is due to Anderson localization. This was the scenario originally studied in \cite{Karpiuk12}, and also the one pertaining to Fig. \ref{twinpeaks}. With respect to Sec. \ref{sectionL}, the localization time $\tau_\text{loc}\equiv \xi^2/D_B$ now plays the role of the Thouless time $\tau_D\equiv L^2/D_B$, and the Heisenberg time $\tau_H=2\pi\overline{\nu} \xi^2$ now refers to a volume of size $\xi$. The case $\xi\ll L$ is extremely interesting since now \emph{the emergence of the CFS peak is a hallmark of localization}. When $\xi\ll L$, no exact solution of the nonlinear $\sigma$-model (\ref{Susy}) is unfortunately available in two dimensions, and consequently there is no exact expression for the CFS contrast for $t>\tau_\text{loc}$. Nevertheless, the time evolution of CFS can be estimated in some limiting cases. First, for $\tau_\text{loc}< t\ll \tau_H$ (i.e. at the onset of the ergodic regime, see Fig. \ref{phase_diagram}), it was suggested by a qualitative argument of renormalization of the diffusion coefficient in the diagrams in Fig. \ref{diagrams} that the CFS peak should increase as $t/\tau_H$ \cite{Karpiuk12}. The isotropic background and the maximum of the CBS peak were on the other hand predicted to remain unchanged at any time $t\gg\tau_B$, namely to be still given by Eqs. (\ref{Background}) and (\ref{CBS_diag}), respectively \cite{Karpiuk12}. As we now show, the CFS contrast for $\xi\ll L$ can also be evaluated in the long-time limit $t\gg\tau_H$ (quantum regime, see Fig. \ref{phase_diagram}). For this purpose, we first recognize that the function $K_E(t)$ in Eq. (\ref{CFS_longt}) that we derived for $L\ll\xi$ is nothing but the so-called form factor: \begin{equation} \label{form_factor} K_E(t)=\intdone{\omega}e^{-i \omega t}K_E(\omega), \end{equation} where $K_E(\omega)=\overline{\delta\nu(E+\omega/2)\delta\nu(E-\omega/2)}/\overline{\nu}(E)^2$ is the Fourier transform of the correlation function of density-of-states fluctuations, which can be rewritten as \cite{Mirlin00} \begin{eqnarray} \label{level_cor} &&K_E(\omega) = -1+\nonumber\\ &&\dfrac{1}{\overline{\nu}(E)^2L^4} \overline{\sum_{i,j}\delta\left(E+\dfrac{\omega}{2}-E_i\right)\delta\left(E-\dfrac{\omega}{2}-E_j\right)}, \end{eqnarray} where the $E_i$ are the energy levels of the disordered system. In fact, the relation (\ref{CFS_longt}) between the CFS peak and the form factor defined by (\ref{level_cor}) turns out to hold very generally, for any ratio of $L$ and $\xi$. This can be explicitly shown by a modal decomposition of the wave function written in momentum space, a task that was accomplished in \cite{Loon14}. With the help of this relation, the problem of calculating the CFS contrast as a function of time boils down to the analysis of the frequency dependence of the correlation function $K_E(\omega)$. In the localization regime $\xi\ll L$ and in the limit of small frequencies, this dependence can be accessed within a simple model of ``correlated localization volumes'', originally introduced by Mott \cite{Mott70}. Let us briefly recall the main lines of this model. We here essentially follow the point of view of \cite{Sivan87, Atland95}: we conceptually divide our 2D disordered system in small patches of volume $\xi^2$, such that in each patch the mean level spacing is $\Delta\equiv(\overline{\nu}\xi^2)^{-1}=2\pi/\tau_H$. Within one patch, two eigenstates experience the usual level repulsion of disordered systems \cite{Mehta} and are therefore far apart in the spectrum. Conversely, let us consider two close levels $E_i$ and $E_j$ such that $|E_i-E_j|\equiv|\omega|\ll\Delta$. The corresponding eigenstates then belong to two distant localization patches. We model the sub-system formed by these two levels by the coupling Hamiltonian \begin{equation} \label{Hmatrix} H_c=\begin{pmatrix} \epsilon_1 & \Delta e^{-|\textbf{r}_1-\textbf{r}_2|/\xi} \\ \Delta e^{-|\textbf{r}_1-\textbf{r}_2|/\xi} & \epsilon_2 \end{pmatrix}. \end{equation} Here $\epsilon_1$ and $\epsilon_2$ are the energy levels in the absence of coupling, and $\Delta e^{-|\textbf{r}_1-\textbf{r}_2|/\xi}$ is the overlap integral between the two uncoupled states, whose wave functions are exponentially localized around $\textbf{r}_1$ and $\textbf{r}_2$, respectively. Due to the coupling, the levels become $\bar\epsilon\pm\sqrt{\delta\epsilon^2/4+\Delta^2e^{-2r/\xi}}$, where $\bar\epsilon\equiv(\epsilon_1+\epsilon_2)/2$, $\delta\epsilon\equiv\epsilon_1-\epsilon_2$ and $r\equiv|\textbf{r}|\equiv|\textbf{r}_1-\textbf{r}_2|$. As in \cite{Sivan87}, we assume that $\bar\epsilon$, $\delta\epsilon$ and $\textbf{r}$ are independent, uniformly distributed random variables, respectively over an interval of size $\Delta$ (for $\bar\epsilon$ and $\delta\epsilon$) and over the volume $L^2$ (for $\textbf{r}$). Performing the integral over $\bar\epsilon$ allows us to get rid of one of the delta functions in Eq. (\ref{level_cor}), which yields \begin{equation} \label{K_eq} K_E(\omega)\sim\int_{-\Delta/2}^{\Delta/2}d\delta\epsilon \int \dfrac{d^2\textbf{r}}{L^2}\ \delta\left(\omega-\sqrt{\delta\epsilon^2+4\Delta e^{-2r/\xi}}\right). \end{equation} In Eq. (\ref{K_eq}), the integral over the difference of localization centers ranges over the full volume of the system, $L^2$. The bounds in the integral over $\delta\epsilon$ account for the fact that the absolute difference between $\epsilon_1$ and $\epsilon_2$ should not be greater that $\Delta$ because we only consider the coupling between states belonging to different localization patches. The two integrals are readily performed, using the inequality $|\omega|\ll\Delta$ and taking the limit $\xi\ll L$. We obtain \begin{equation} \label{Komega_final} K_E(\omega)\sim\left(\dfrac{\xi}{L}\right)^2\ln^2\left(\dfrac{|\omega|}{2\Delta}\right),\ |\omega|\ll \Delta. \end{equation} Note that in deriving Eq. (\ref{Komega_final}), we implicitly assumed $\omega\ne0$. In order to describe long times, we must also include the contribution $\omega=0$, which comes from the diagonal terms $i=j$ in the sum in Eq. (\ref{level_cor}). These terms yield the contribution $\delta(\omega)/[\overline{\nu}(E) L^2]$, which describes the self correlation of one energy level. Adding it to Eq. (\ref{Komega_final}) and performing the Fourier transform with respect to $\omega$, we obtain the final form of $K_E(t)$ and of the CFS contribution to the momentum distribution at $\bs{k}=+\bs{k}_0$, at long times $t\gg\tau_H$: \begin{equation} \label{CFS_longt_loc} \av{n}_\text{CFS}(\bs{k}_0,t)\simeq\intdone{E}\dfrac{A^2(\bs{k}_0,E)}{2\pi\overline{\nu}(E)}\left[1-\alpha\dfrac{\ln(\beta t/\tau_H)}{t/\tau_H}\right]. \end{equation} In writing Eq. (\ref{CFS_longt_loc}), we have introduced two phenomenological parameters, $\alpha$ and $\beta$, whose precise determination is not accessible from the present approach. $\alpha$ accounts for the fact that in a real system, the distributions of $\bar\epsilon$, $\delta\epsilon$ and $r$ may not be exactly uniform, while $\beta$ accounts for the fact that the strength of the coupling terms in the Hamiltonian (\ref{Hmatrix}) may slightly differ from $\Delta$. The appearance of a logarithm in Eq. (\ref{CFS_longt_loc}) is however typical of 2D disordered systems \cite{Atland95}. This has to be contrasted with one-dimensional or quasi one-dimensional geometries for which $K_E(\omega)\propto\ln[|\omega|/(2\Delta)]$, leading eventually to a purely algebraic decay $\propto(t/\tau_H)^{-1}$ of the second term in the right-hand-side of Eq. (\ref{CFS_longt_loc}) \cite{Micklitz14, Loon14}. \subsection{Numerical simulations for $\xi\ll L$} \label{numerics_loc} Let us now confront the predictions (\ref{Background}), (\ref{CBS_diag}) and (\ref{CFS_longt_loc}) with numerical simulations. To describe the localization regime, it is necessary to make the localization length smaller than the system size, and it is thus mandatory to use a stronger value of the disorder amplitude $V_0$. The main panel of Fig. \ref{SpectralF5} displays the numerically computed spectral function $A(\bs{k}_0,E)$ for $V_0=5$ and $k_0=1.5$. In sharp contrast with Sec. \ref{numerics_WD} where disorder was relatively weak, the spectral function has now a maximum at negative energy and a long tail toward high energies. This raises two new problems. First, for a given system size $L$ it is hard to fulfill the inequality $\xi(E)\ll L$ for all energies. For instance, when $L=100\pi$, the energy $E_L$ such that $\xi(E_L)=L$ is slightly above zero, thus distinctly above the maximum of the spectral function, see the main panel of Fig. \ref{SpectralF5}. Therefore, many particles have $E>E_L$ and thus have a localization length $\xi(E)> L$. Second, unlike in Sec. \ref{numerics_WD} the Heisenberg time (which is proportional to the square of the localization length) now varies much faster with $E$ than the spectral function. As a consequence, it is not even possible to identify a single Heisenberg time for localized atoms. In order to nevertheless consider a ``clean'' situation where all atoms are localized with approximately the same Heisenberg time, we introduce a filtering in energy in the time-propagation algorithm: at $t=0$ we apply the operator $\exp[-(\hat{H}-\epsilon_0)^2/(2\Delta\epsilon^2)]$ to the initial state $|\bs{k}_0\rangle$, and only then propagate it with the evolution operator $\exp(-i\hat{H}t)$. With this procedure the energies $E$ involved in transport roughly lie in the interval $[\epsilon_0-\Delta\epsilon/2,\epsilon_0+\Delta\epsilon/2]$, which is chosen so that $\xi(E)\ll L$ for all $E$ within that interval. In the rest of this section, we choose $\epsilon_0=-2.5$ and $\Delta\epsilon=0.4$ for $V_0=5$ and $k_0=1.5$, see Fig. \ref{SpectralF5}. With these parameters, the localization length given by the transfer-matrix approach at $E=\epsilon_0$ is $\xi\simeq 3.9$. From a theoretical point of view, the filtering in energy amounts to performing the replacement $A(\bs{k}, E)\rightarrow A(\bs{k}, E)\exp\left[-(E-\epsilon_0)^2/(2\Delta\epsilon^2)\right]$ (for both $\bs{k}$ and $\bs{k}_0$) in all formulas. \begin{figure} \includegraphics[width=1\linewidth]{SpectralF5.eps} \caption{(Color online) Spectral function $A(\bs{k}_0,E)$ as a function of energy, for $V_0=5$ and $k_0=1.5$ (lengths, momenta and energies are respectively given in units of $\zeta$, $\zeta^{-1}$ and $1/(m\zeta^2)$, where $\zeta$ is the correlation length of the random potential). For the time evolution, only energies in the narrow band $[\epsilon_0-\Delta\epsilon/2,\epsilon_0+\Delta\epsilon/2]$ are selected, where $\epsilon_0=-2.5$ and $\Delta\epsilon=0.4$. The inset shows the localization length given by the transfer-matrix approach as a function of energy, for the same value of $V_0$. Data are averaged over $16$ disorder realizations.} \label{SpectralF5} \end{figure} \begin{figure} \includegraphics[width=0.9\linewidth]{background_V5.eps} \caption{(Color online) Cut along $k_x=0$ of the momentum distribution, describing the radial shape of its isotropic, ring-shaped part. Points were obtained from numerical simulations of plane-wave propagation in a 2D speckle potential in the limit $\xi\ll L$ ($V_0=5$, $k_0=1.5$, $\epsilon_0=-2.5$, $\Delta\epsilon=0.4$), after averaging over $240$ disorder realizations. $L$ is set to $100\pi$. The red curve is the prediction (\ref{Background}), in which spectral functions and density of states are computed numerically. There is no adjustable parameter.} \label{background_v5} \end{figure} We first show in Fig. \ref{background_v5} a radial cut along $k_x=0$ of the numerically computed momentum distribution (black points). These points are expected to describe the radial shape of the isotropic background. The chosen time is $t=1500$, i.e. well beyond the transport mean free time which is $\tau_B\simeq 2.6$. In the same plot we show the theoretical prediction (\ref{Background}) fed with the numerically computed spectral functions $A(\bs{k}_0,E)$, $A(\bs{k},E)$, and density of states $\av{\nu}(E)$. The agreement is very good and may come as a surprise if we remember that Eq. (\ref{Background}) has been actually derived in the weak-disorder limit. It demonstrates the general validity of Eq. (\ref{Background}) for the isotropic background, even in the deep localization regime $\xi\ll L$ and for rather strong disorder, provided the \emph{exact} $A(\bs{k},E)$ and $\av{\nu}(E)$ are used in the computation. This suggests that at strong disorder, all interference corrections to the isotropic background boil down to a renormalization of the scattering mean free path while the global topology of the diagram in Fig. \ref{diagrams}(a) remains valid. Note in passing that the dip around $k_y=0$ is less pronounced in Fig. \ref{background_v5} than in Fig. \ref{background}. This is a direct consequence of the long energy tail of the spectral function at stronger disorder. \begin{figure}[h] \centering\includegraphics[width=1\linewidth]{CBSCFS_strong.eps} \caption{(Color online) \label{CBSCFS_strong} Main panel: contrast of the CBS (upper green symbols) and CFS (lower red symbols) peaks as a function of $t/\tau_H$, obtained from numerical simulations of plane-wave propagation in the localization regime $\xi\ll L$, for the same parameters as in Fig. \ref{background_v5} (here $\tau_H\simeq40$ \cite{footnote2}). Circles were obtained for $L=50\pi$, squares for $L=80\pi$ and crosses for $L=100\pi$. Data are averaged over $240-1600$ disorder realizations (depending on the value of $L$) and over a time window $\Delta t=40$. The dashed and solid curves are the theoretical predictions (\ref{Contrast_CBS_loc}) and (\ref{Contrast_CFS_loc}), respectively. Inset: contrasts at short times $t\ll\tau_H$. For comparison, the scaled background $\overline{n}_D(k_0,t)/\overline{n}_D(k_0,\infty)$ is also shown (blue crosses). Both the background and the CBS contrast become time independent after a few $\tau_B$, while the CFS contrast increases slowly, linearly in time, as highlighted by the dotted line. } \end{figure} The CBS contrast for $\xi\ll L$ is still given by the ratio of Eqs. (\ref{Background}) and (\ref{CBS_diag}), and has thus the same expression as in Sec. \ref{sectionL}: \begin{equation} \label{Contrast_CBS_loc} C_\text{CBS}=1,\ t\gg\tau_B. \end{equation} On the other hand, the CFS contrast at times $t\gg\tau_H$ now follows from Eq. (\ref{CFS_longt_loc}): \begin{equation} \label{Contrast_CFS_loc} C_\text{CFS}=1-\alpha\dfrac{\ln(\beta t/\tau_H)}{t/\tau_H},\ t\gg\tau_H. \end{equation} These two relations are shown in the main panel of Fig. \ref{CBSCFS_strong} (dashed and solid curves, respectively) together with the CBS and CFS contrasts obtained from our numerical simulations (green and red symbols, respectively), for $\tau_H\simeq 40$ \cite{footnote2}. Circles were obtained for a system size $L=50\pi$, squares for $L=80\pi$ and crosses for $L=100\pi$. We find that Eq. (\ref{Contrast_CFS_loc}) well reproduces the numerical results for $\alpha=0.5\pm 0.1$ and $\beta=2.3\pm 0.1$ [for the fit we only consider times larger than $\tau_H$, which is the limit of validity of Eq. (\ref{CFS_longt_loc})]. Note that as opposed to the case $L\ll\xi$ (see Fig. \ref{CBSCFS_weak}), the contrast of the CFS no longer depends on $L$, as expected in the localization regime. The inset additionally shows the CBS and CFS contrasts together with the background contribution to the momentum distribution at short times $t\ll\tau_H$. The observed behavior is qualitatively the same as in the case $L\ll\xi$: both the background and CBS contrast become time independent after a few $\tau_B$, while the CFS contrast increases slowly in time. This increase is compatible with the linear law estimated in \cite{Karpiuk12}, though the latter is seen in a rather small time interval. In order to have a better estimation of the validity of the theoretical expression for the CFS contrast at long times, Eq. (\ref{Contrast_CFS_loc}), we replot in Fig. \ref{CFS_long} the quantity $(1-C_\text{CFS})t/\tau_H$ as a function of $t/\tau_H$. In this representation, the numerical points increase logarithmically in time, in full agreement with Eq. (\ref{Contrast_CFS_loc}). \begin{figure} \includegraphics[width=0.95\linewidth]{CFS_long.eps} \caption{(Color online) \label{CFS_long} Red crosses: numerical values of $(1-C_\text{CFS})t/\tau_H$ plotted as a function of $t/\tau_H$, for the same parameters as in Fig. \ref{CBSCFS_strong}, with $L$ set to $100\pi$. The dashed curve is the function $\alpha\ln(\beta x)$, where $\alpha=0.5$ and $\beta=2.3$. Data are averaged over $240$ disorder realizations and over a time window $\Delta t=40$. } \end{figure} For the sake of completeness, we finally show in Fig. \ref{CFS_width} a plot of the width at half maximum in momentum space of both the numerical CBS and CFS peaks as a function of time. As for the contrast, we see that the evolutions of the two peaks are different. Over a time scale of the order of the mean free time, the CBS width quickly converges to a value of the order of a few $\xi^{-1}$. The CFS width, on the other hand, slowly decreases, until it reaches the same value as the CBS width after a few Heisenberg times, suggesting identical CBS and CFS profiles at very long times. \begin{figure}[h] \includegraphics[width=0.9\linewidth]{CFS_CBS_width.eps} \caption{(Color online) \label{CFS_width} Width of the CBS and CFS peaks in momentum space as a function of $t/\tau_H$, in the localization regime $\xi\ll L$, for the same parameters as in Fig. \ref{CBSCFS_strong}. $L$ is set to $100\pi$. At long times, both widths converge to the same constant value, which is of the order of a few $\xi^{-1}$. Data are averaged over $240$ disorder realizations and over a time window $\Delta t=40$. } \end{figure} \section{Experimental scenario} \label{feasibility} In Sec. \ref{sectionxi}, we showed that when $\xi\ll L$ the CFS peak is a signature of Anderson localization, which confines atoms in a region of size $\xi$. However, we considered an ideal scenario where: (i) the dynamics is supported by a single energy $E=\epsilon_0$ and (ii) the initial state is a plane wave. In current state-of-the-art experiments on ultracold atoms however, these two conditions are not fulfilled. Indeed, all the energy components authorized by the spectral function shown in Fig. \ref{SpectralF5} contribute to transport, which means in particular that both diffusive and localized atoms are present. In addition, the initial state is never a plane wave but rather a wave packet of finite size $(\Delta k)^{-1}\ne\infty$ in configuration space. It thus remains important to clarify whether, within this non-ideal scenario, the CFS peak due to localized atoms is visible or not. This is the object of the present section. \subsection{Effect of a broad energy distribution} Let us first address the effect of a broad energy distribution on the CFS dynamics. For this purpose, as in Sec. \ref{numerics_loc} we numerically carry out the time evolution of a plane wave [$(\Delta k)^{-1}=\infty$] for $V_0=5$ and $k_0=1.5$, but this time \emph{without} applying the filtering in energy, \begin{figure}[h] \includegraphics[width=0.95\linewidth]{CFS_L_effect.eps} \caption{(Color online) \label{CFS_L_effect} Colored symbols: contrast of the CFS peak as a function of time, obtained from numerical simulations of plane-wave propagation [$(\Delta k)^{-1}=\infty$] for $V_0=5$ and $k_0=1.5$, without using any filtering in energy (lengths, momenta, energies and times are respectively given in units of $\zeta$, $\zeta^{-1}$, $1/(m\zeta^2)$ and $m\zeta^2$, where $\zeta$ is the correlation length of the random potential). The four curves correspond to different values of $L$. Data are averaged over $400$ disorder realizations and over a time window $\Delta t=40$. For comparison, the dashed curve shows the theoretical prediction (\ref{Contrast_CFS_loc}), corresponding to the ideal case of a single energy component $E=\epsilon_0=-2.5$, for which $\tau_H\simeq 40$. } \end{figure} such that now both localized and diffusive atoms coexist. The contrast of the CFS peak obtained in this way is shown in Fig. \ref{CFS_L_effect} as a function of time, for four values of the system size $L$. Several observations can be made. First, when $L$ is small, the CFS contrast decreases with $L$. This effect is due to diffusive atoms, which fulfill $L<\xi(E)$ and thus produce a CFS peak because of their confinement in the volume $L^2$ (mechanism discussed in Sec. \ref{sectionL}). Second, for the parameters used in Fig. \ref{CFS_L_effect}, the CFS contrast no longer visually changes with $L$ when $L\gtrsim 50\pi$. This means that in this limit -- which effectively corresponds to the experimental scenario $L=\infty$ -- the observed CFS peak is \emph{entirely due to localized atoms} (at smaller $L$ this is also the case at short enough times). Overall, the CFS contrast is however smaller than the ideal situation of Sec. \ref{sectionxi} where the dynamics was supported by a single, localized energy [Eq. (\ref{Contrast_CFS_loc}), dashed curve in Fig. \ref{CFS_L_effect}], because at a given time $t$ many localized atoms have an Heisenberg time $\tau_H(E)>t$ and thus have not yet contributed to the CFS peak. \subsection{Effect of the size of the wave packet} Having discussed the effect of a broad energy distribution, we now additionally consider the effect of the finite size of the initial wave packet. We show in Fig. \ref{CFS_Deltak_effect} the contrast of the CFS peak as a function of time for $L=100\pi$, obtained from numerical simulations starting from a Gaussian wave packet of width $(\Delta k)^{-1}\ne\infty$ rather than from a plane wave (as before $V_0=5$ and $k_0=1.5$ and no filtering in energy is applied). The finite value of $(\Delta k)^{-1}$ leads to a decay of the CFS contrast, well visible in the figure. This phenomenon can be traced back to Eq. (\ref{rhokprimet}): using a Gaussian wave packet amounts to taking $n_0(\bs{k}')=(4\pi/\Delta k^2)\exp[-(\bs{k}'-\bs{k}_0)^2/(\Delta k^2)]$ instead of $(2\pi)^2\delta(\bs{k}'-\bs{k}_0)$, and thus to convolving the CFS peak obtained for a plane wave with a Gaussian function. From a physical point of view, the finite size of the wave packet cuts multiple scattering trajectories whose start and end points are separated by more than $(\Delta k)^{-1}$, as for the CBS effect \cite{Cherroret12}. \begin{figure}[h] \includegraphics[width=0.95\linewidth]{CFS_Deltak_effect.eps} \caption{(Color online) \label{CFS_Deltak_effect} Colored symbols: contrast of the CFS peak as a function of time, obtained from numerical simulations starting from a Gaussian wave packet of finite width $(\Delta k)^{-1}$, for $V_0=5$, $k_0=1.5$ and $L=100\pi$, without using any filtering in energy. The three curves correspond to different values of $(\Delta k)^{-1}$, increasing from bottom to top. Data are averaged over $400$ disorder realizations and over a time window $\Delta t=40$. For comparison, the dashed curve shows the theoretical prediction (\ref{Contrast_CFS_loc}), corresponding to the ideal case of a single energy component $E=\epsilon_0=-2.5$, for which $\tau_H\simeq 40$. } \end{figure} \section{Summary and concluding remarks} \label{Conclusion} This paper was devoted to a systematic study of the momentum distribution of a matter wave launched with finite velocity in a 2D random, speckle potential. In particular, we analyzed in detail the slowly evolving coherent forward scattering peak arising in the momentum distribution at long enough times. We showed that the emergence of this peak is conditional on the presence of a mechanism of confinement which allows to enhance the interference mechanism scattering particles in the forward direction. This confinement can arise in the situation where particles propagate diffusively in a bounded volume, or because of Anderson localization, which prevents transport beyond scales of the order of the localization length $\xi$. We studied both numerically and theoretically the two limits $L\ll\xi$ and $\xi\ll L$, for which we summarize the asymptotic expressions for the CFS contrast in Table \ref{table1}, for the ideal case where transport is supported by a single energy component. \begin{table}[h] \begin{tabular}{>{\centering}p{3.7cm}|>{\centering}p{2.2cm}|>{\centering}p{2.3cm}} time & $L\ll\xi$ $\tau_H\equiv 2\pi\overline{\nu}L^2$ & $\xi\ll L$ $\tau_H\equiv 2\pi\overline{\nu}\xi^2$ \tabularnewline \hline\hline Diffusive regime $\tau_B\ll t\ll\text{min}(L,\xi)^2/D_B$ & $\propto \dfrac{1}{k_0\ell_B}$& $\propto \dfrac{1}{k_0\ell_B}$ \tabularnewline \hline Ergodic regime $\text{min}(L,\xi)^2/D_B\ll t\ll\tau_H$ & $2\dfrac{t}{\tau_H}$ $\propto\dfrac{t}{\tau_H}$ \tabularnewline \hline Quantum regime\\ $t\gg\tau_H$ & $1-\dfrac{1}{12 (t/\tau_H)^2}$& $1-\alpha\dfrac{\ln(\beta t/\tau_H)}{t/\tau_H}$ \tabularnewline \end{tabular} \caption{\label{table1} Summary of asymptotic expressions for the CFS contrast $C_\text{CFS}$ in a 2D disordered system. $\alpha=0.5\pm 0.1$ and $\beta=2.3\pm 0.1$.} \end{table} From our results, it thus turns out that CFS of a matter wave could be experimentally observed by either artificially confining atoms in some finite regime of space, for instance with an optical potential with steep enough edges, or, more interestingly, by achieving Anderson localization. In current experimental setups, for instance that in \cite{Josse12}, atomic motion is not limited by any artificial boundary in the relevant directions of propagation, which corresponds to $L=\infty$. Consequently, any observation of the CFS effect using those setups \emph{would be a genuine signature of Anderson localization}. In an experiment, the visibility of CFS can be reduced because the matter wave supports many energy components and has initially a finite spatial width, but we showed that this reduction of visibility is rather small. From a theoretical point of view, we saw that, except for short times, the dynamics of the CFS peak is not accessible from perturbation theory and requires the help of the nonlinear $\sigma$-model or of another non-perturbative approach. Still, in two dimensions there is presently no exact solution of the $\sigma$-model for $\xi\ll L$, such that no exact expression of the CFS peak is available at all times in this limit. This conclusion also applies to the three-dimensional case where the Anderson transition is present, which offers an interesting theoretical challenge for future works. In the search for the CFS peak, ultracold atoms have many advantages, including the possibility for \emph{in situ} measurements of the velocity distribution. In principle however, the CFS peak could be also observed with other types of waves and in particular with classical waves propagating in disordered media. In this context, experiments often involve a ``scattering setup'' in which the wave is sent from outside the disordered system and transport properties are probed in transmission or reflection. While the reflection is well known to exhibit a prominent CBS peak \cite{Bayer93, Tourin97, Wolf85, Albada85}, the possibility of observing a CFS peak in the transmitted profile is more speculative. Indeed, in this setup both a high signal-to-noise ratio and a good angular resolution would be required to detect the CFS peak, the latter being very narrow and sitting on top of an exponentially small transmission signal. Closer to the situation described in the present work on the other hand, the scenario of transverse localization of light in paraxial geometries seems more promising \cite{Raedt89, Schwartz07}. Coming back to the atomic context, it would finally be interesting to see how the CBS and CFS dynamics are perturbed by the presence of weak interactions between atoms, typically described by the nonlinear Gross-Pitaevskii equation for bosons. For a wave packet expanding in disorder, a weak nonlinearity is known to partially destroy Anderson localization and to restore a transport slower than diffusion \cite{Cherroret14, Kopidakis08, Pikovsky08, Basko11}. Nonlinearity-driven subdiffusion could manifest as well in momentum space and alter the CBS and CFS peaks, as what is known to happen to CBS in stationary setups \cite{Hartung08}. \section*{Acknowledgements} NC thanks Cord M\"uller for useful advice and comments, and the hospitality of the CQT, where part of this work was carried out. SG, CM and DD thank the Institut Fran\c{c}ais de Singapour (French Ministry of Foreign Affairs) and the National University of Singapore for supporting this work through the Merlion Programme. The authors acknowledge financial support from the French agency ANR (Project No. 11-B504-0003 LAKRIDI). This work was granted access to the HPC resources of TGCC under the allocation 2014-057083 made by GENCI (Grand Equipement National de Calcul Intensif) and to the HPC resources of The Institute for scientific Computing and Simulation financed by Region Ile de France and the project Equip@Meso (reference ANR-10-EQPX-29-01). The Centre for Quantum Technologies is a Research Centre of Excellence founded by the Ministry of Education and the National Research Foundation.
\section{Introduction} A quantized vortex in a superfluid is a topological defect, and it reflects the symmetry of the macroscopic wave function. For example, a single-component Bose--Einstein condensate (BEC) is described by a complex wave function $\psi \propto e^{i \phi}$ that has a phase degree of freedom $\phi$, and its symmetry group is U(1). As a result, a vortex is characterized by an integer winding number, known as the Onsager--Feynman quantization~\cite{Onsager,Feynman}. In a BEC of atoms with spin degrees of freedom, the ground-state manifold has a more complicated topology, and a variety of spin vortices exist, such as a half-quantum vortex~\cite{Leonhardt} and a polar-core vortex~\cite{Sadler}. Topological properties of spinor BECs have been studied by many researchers~\cite{Choi,Ray,Ohmi,Ho,Khawaja,Savage,Ruo,Makela,Semenoff,Barnett,Kawaguchi,Michikazu,Pietila,Borgh,Kobayashi}. The ground-state magnetic phase of a spin-1 BEC is the ferromagnetic or polar phase depending on the interaction coefficients~\cite{Stenger}. The symmetry group $G_f$ of the ferromagnetic phase is SO(3)~\cite{Ho}, i.e., there is a one-to-one correspondence between a state in the ferromagnetic phase and an element in $G_f$. The change in the spin state around a core of a spin vortex is therefore equivalent to a closed loop in the $G_f$ manifold, which is classified by the fundamental group $\pi_1(G_f) = \mathbb{Z}_2$. For the polar phase, the symmetry group $G_p$ is ${\rm U}(1) \times S^2 / \mathbb{Z}_2$, and its fundamental group is $\pi_1(G_p) = \mathbb{Z}$~\cite{Makela}. Let us consider a situation in which the magnetic phase containing a spin vortex changes (spatially or temporally) from the ferromagnetic to polar phases, or vice versa. Since a change in the magnetic phase is accompanied by a change in its symmetry group between $G_f$ and $G_p$, the topology of a spin vortex is forced to change between elements in $\pi_1(G_f)$ and $\pi_1(G_p)$. Such situations have been considered by several researchers. In Ref.~\cite{Borgh}, connections of spin-vortex lines lying on both sides of the two magnetic phases in a rotating spin-1 BEC were studied. Tracing a vortex line from one side of the magnetic phase, it undergoes various changes across the interface between the magnetic phases. In Ref.~\cite{Kobayashi}, structures of spin-vortex cores were studied. This problem is also regarded as the spin-vortex connection between different phases, since the symmetry group of the spin state changes radially from infinity to the vortex core. In Ref.~\cite{Hoshi}, the sudden change in the interaction parameter from polar to ferromagnetic was examined; a half-quantum vortex in the polar phase was found to magnetize breaking the rotational symmetry. In the present paper, we investigate the dynamics of a vortex dipole (a vortex-antivortex pair) in a spin-1 BEC, where the spin-dependent interaction parameter is spatially distributed, and the ferromagnetic and polar regions are separated by a phase boundary. We create a vortex dipole on one side of the magnetic regions, and it moves toward the phase boundary. When a vortex dipole has a sufficiently large velocity, it can penetrate the phase boundary~\cite{Aioi}. Passing through the phase boundary, the vortices experience a change in the magnetic phase, and consequently, the topological properties of the vortices are forced to change. We will show that the system exhibits a rich variety of spin dynamics that reflect the symmetry group of each magnetic phase. When a singly-quantized vortex dipole moves from the ferromagnetic to the polar phase, it transforms into a spin vortex dipole in which two ferromagnetic cores exhibit a leapfrogging behavior, or it transforms into a half-quantum vortex dipole. When a singly-quantized vortex dipole moves from the polar to the ferromagnetic phase, it transforms to a spin vortex dipole with ferromagnetic cores surrounded by half-quantum vortices. This paper is organized as follows. Section~\ref{s:formulation} formulates the problem and provides a numerical method. Section~\ref{s:f_to_p} shows the dynamics of vortex dipoles traveling from ferromagnetic to polar phases, and Sec.~\ref{s:p_to_f} shows those for traveling from polar to ferromagnetic phases. Section~\ref{s:conc} presents our conclusions from this study. \section{Formulation of the problem} \label{s:formulation} We use mean-field theory to analyze a BEC of spin-1 atoms. The macroscopic wave functions $\psi_m(\bm{r}, t)$ for magnetic sublevels $m = 0, \pm1$ obey the Gross--Pitaevskii equation, \begin{subequations} \label{GP} \begin{eqnarray} i\hbar \frac{\partial \psi_0}{\partial t} & = & -\frac{\hbar^2}{2M} \nabla^2 \psi_0 + g_0 \rho \psi_0 + \frac{g_1}{\sqrt{2}} \left( F_+ \psi_1 + F_- \psi_{-1} \right), \nonumber \\ & & \\ i\hbar \frac{\partial \psi_{\pm 1}}{\partial t} & = & -\frac{\hbar^2}{2M} \nabla^2 \psi_{\pm 1} + g_0 \rho \psi_{\pm 1} \nonumber \\ & & + g_1 \left( \frac{1}{\sqrt{2}} F_{\mp} \psi_0 \pm F_z \psi_{\pm 1} \right), \end{eqnarray} \end{subequations} where $M$ is the atomic mass and $\rho = |\psi_1|^2 + |\psi_0|^2 + |\psi_{-1}|^2$ is the atomic density. The spin densities in Eq.~(\ref{GP}) are defined as $\bm{F} = \sum_{mm'} \psi_m^* \bm{f}_{mm'} \psi_{m'}$, where $\bm{f}$ is the vector of spin-1 matrices, and $F_\pm = F_x \pm i F_y$. The interaction coefficients in Eq.~(\ref{GP}) are given by $g_0 = 4 \pi \hbar^2 (a_0 + 2 a_2) / (3M)$ and $g_1 = 4 \pi \hbar^2 (a_2 - a_0) / (3M)$, where $a_0$ and $a_2$ are the $s$-wave scattering lengths with colliding channels with total spin 0 and 2, respectively. For simplicity, we focus on an infinite uniform system and Eq.~(\ref{GP}) contains no external potential terms. We assume that the quadratic Zeeman effect is negligible. The ground-state magnetic phase depends on the sign of the interaction coefficient $g_1$. For $g_1 < 0$, the spin-dependent interaction favors the ferromagnetic ground-state. Because of the U(1) gauge and spin rotation symmetries, the ground-state manifold of the ferromagnetic state is expressed as~\cite{Ho} \begin{equation} \label{ferro} \zeta = e^{i\phi} R(\alpha, \beta, \gamma) \left( \begin{array}{c} 1 \\ 0 \\ 0 \end{array} \right) = e^{i(\phi - \gamma)} \left( \begin{array}{c} e^{-i\alpha} \cos^2 \frac{\beta}{2} \\ \sqrt{2} \sin \frac{\beta}{2} \cos \frac{\beta}{2} \\ e^{i\alpha} \sin^2 \frac{\beta}{2} \end{array} \right), \end{equation} where $e^{i\phi}$ and $R(\alpha, \beta, \gamma) = e^{-i f_z \alpha} e^{-i f_y \beta} e^{-i f_z \gamma}$ are the U(1) and SO(3) rotations, respectively. Since $\phi$ can be absorbed into $\gamma$ in Eq.~(\ref{ferro}), the symmetry group of the ferromagnetic state is $G_f =$ SO(3). For $g_1 > 0$, the ground state is the polar phase given by \begin{equation} \label{polar} \zeta = e^{i\phi} R(\alpha, \beta, \gamma) \left( \begin{array}{c} 0 \\ 1 \\ 0 \end{array} \right) = e^{i\phi} \left( \begin{array}{c} -\frac{1}{\sqrt{2}} e^{-i\alpha} \sin \beta \\ \cos\beta \\ \frac{1}{\sqrt{2}} e^{i\alpha} \sin \beta \end{array} \right). \end{equation} Since the state in Eq.~(\ref{polar}) is independent of $\gamma$ and invariant with respect to $\alpha \rightarrow \alpha + \pi$, $\beta \rightarrow \pi - \beta$, and $\phi \rightarrow \phi + \pi$, the symmetry group of the polar state is $G_p = {\rm U}(1) \times S^2 / \mathbb{Z}_2$. We consider a situation in which the interaction coefficient $g_1$ is spatially distributed as \begin{equation} \label{g2} g_1(\bm{r}) = \left\{ \begin{array}{cc} g_f < 0 & (x < 0), \\ g_p > 0 & (x > 0), \end{array} \right. \end{equation} where $g_f$ and $g_p$ are negative and positive constants, respectively. In the following, we will take $g_f = -0.01 g_0$ and $g_p = 0.01 g_0$. Such a space-dependent interaction coefficient may be realized by, e.g., an optical Feshbach technique. For the interaction coefficient $g_1$ in Eq.~(\ref{g2}), the ground state is the ferromagnetic state for $x \rightarrow -\infty$ and the polar state for $x \rightarrow \infty$. Their phase boundary is located at $x \simeq 0$, where the two phases are smoothly connected over the spin healing length. When $\zeta|_{x \rightarrow -\infty} = (1, 0, 0)^T$, the spin state for $x \rightarrow \infty$ must be $\zeta|_{x \rightarrow \infty} = (1, 0, 1)^T / \sqrt{2}$ in the ground state, where the superscript $T$ stands for the transpose. Connection to other polar states, such as $\zeta|_{x \rightarrow \infty} = (0, 1, 0)^T$, does not minimize the energy. Spin vortices in a magnetic phase with symmetry group $G$ are classified by its fundamental group $\pi_1(G)$. For the ferromagnetic phase, the fundamental group of the symmetry group is $\pi_1(G_f) = \mathbb{Z}_2$, and there are only two topological states: no-vortex and vortex states. An expression of the vortex state far from the core located at the origin is given by $\zeta = (e^{i \theta}, 0, 0)^T$, where $\theta = {\rm arg} (x + i y)$. For the polar phase, $\pi_1(G_p) = \mathbb{Z}$, and corresponding vortex states are written as $\zeta = (e^{i \theta}, 0, e^{i n \theta})^T / \sqrt{2}$ with an integer $n$; this is a singly-quantized vortex for $n = 1$ and a half-quantum vortex for $n = 0$. The topological properties of spin vortices are thus different in the regions $x < 0$ and $x > 0$ for the inhomogeneous interaction coefficient $g_1$ in Eq.~(\ref{g2}). In the following, we study the dynamics of a vortex dipole traveling from $x < 0$ to $x > 0$, or vice versa. We expect that the topology of the vortex dipole changes dynamically at the interface between the ferromagnetic and polar phases. We numerically solve Eq.~(\ref{GP}) using the pseudo-spectral method~\cite{Recipes}. The initial state is the ground state for the inhomogeneous interaction coefficient $g_1$ in Eq.~(\ref{g2}). The ground state is numerically obtained by the imaginary-time propagation method, in which the $i$ on the left-hand side of Eq.~(\ref{GP}) is replaced by $-1$. In numerical simulations, a small numerical noise is added to the initial state to break artificial symmetry. A periodic boundary condition is imposed by the pseudo-spectral method. The size of the system is taken to be sufficiently large, and the boundary condition does not affect the dynamics of the vortices. \section{Propagation of a vortex dipole from ferromagnetic to polar phase} \label{s:f_to_p} First we examine the dynamics of a vortex dipole traveling from the ferromagnetic phase to the polar phase. We prepare the ground state of Eq.~(\ref{g2}), where the atomic density far from the phase boundary is $\rho_0$. The spin state is $\zeta = (1, 0, 0)^T$ for $x \ll \xi_s$, and $\zeta = (1, 0, 1)^T / \sqrt{2}$ for $x \gg \xi_s$, where $\xi_s = \hbar / (M g_1 \rho_0)^{1/2}$ is the spin healing length. The spin healing length and the spin-wave velocity $v_s = (g_1 \rho_0 / M)^{1/2}$ give the characteristic size and velocity of the spin dynamics, which define the characteristic time scale $\tau = \xi_s / v_s$. In the numerical simulations, a vortex dipole is created by imprinting the phase on the wave functions as \begin{equation} \label{phase} \psi_m(z) \rightarrow \frac{(z - z_+) (z - z_-)^*}{|(z - z_+) (z - z_-)|} \psi_m(z), \end{equation} on each numerical grid, where $z = x + i y$, and $z_\pm$ are the positions of a vortex and an antivortex. In the present initial state, we imprint the vortices at $z_\pm = -5 \xi_s \pm i \Delta / 2$, where $\Delta$ is the distance between the vortex and the antivortex. The vortex dipole created on the ferromagnetic side ($x < 0$) then moves in the $+x$ direction at a velocity $\simeq \hbar / (M \Delta)$. \begin{figure}[tbp] \includegraphics[width=8cm]{fig1.eps} \caption{ (Color online) Time evolution of a vortex dipole traveling from the ferromagnetic phase to the polar phase, where the initial distance between the vortices is $\Delta = 1.6\xi_s$. (a)--(f) show $|\psi_1|^2$, $|\psi_{-1}|^2$, and $\sum_m |\psi_m|^2$, and (g) shows $|\psi_1|^2$ and $|\psi_{-1}|^2$. The density $|\psi_0|^2$ is always negligible. The arrows in (a) indicate the directions in which the vortices circulate. The windows of each panel are (a)--(e) $-4 \xi_s < x < 4 \xi_s$ and $-3 \xi_s < y < 3 \xi_s$, (f)$10 \xi_s < x < 18 \xi_s$ and $-3 \xi_s < y < 3 \xi_s$, and (g) $2 \xi_s < x < 4 \xi_s$ and $-2 \xi_s < y < 2 \xi_s$. Note that the scale of (f) is the same as it is for (a)--(e). See the Supplemental Material for a movie of the dynamics of $|\psi_1|^2$. } \label{f:f_to_p1} \end{figure} Figure~\ref{f:f_to_p1} shows the dynamics of a vortex dipole for $\Delta = 1.6 \xi_s$. As the vortex dipole approaches the region of $x \simeq 0$, the phase boundary is deformed, as shown in Fig.~\ref{f:f_to_p1}(b). Passing through the phase boundary, the vortex dipole generates a complicated spin texture, as shown in Figs.~\ref{f:f_to_p1}(c) and \ref{f:f_to_p1}(d), and transforms into a spin vortex dipole traveling in the polar phase, as shown in Fig.~\ref{f:f_to_p1}(e). Despite the formation of the spin texture, the total density is almost homogeneous except for that of the vortex cores. Figure~\ref{f:f_to_p1}(g) shows in detail the vortex dynamics on the polar side. The cores of the vortex-antivortex pair in the $m = \pm 1$ components are occupied by the $m = \mp 1$ components, and the vortex dipoles in both components rotate around one another in a leapfrogging manner. We note that this dynamics is similar to the well-known leapfrogging behavior of vortices~\cite{Helmholtz}, but their mechanisms are quite different; this will be discussed below. As the vortex dipole travels in the polar phase, its cores gradually expand, and the total density gradually increases, as shown in Fig.~\ref{f:f_to_p1}(f). This relaxation occurs as the leapfrogging behavior causes spin wave radiation (see a movie in the Supplemental Material), dissipating the energy of the vortex dipole. \begin{figure}[tbp] \includegraphics[width=8cm]{fig2.eps} \caption{ (Color online) Spatiotemporal image of the magnetization density $|\bm{F}|$ during $10 < t / \tau < 16$. The holes indicated by the arrows correspond to the vortex cores in Fig.~\ref{f:f_to_p1}(b). The parameters are the same as those in Fig.~\ref{f:f_to_p1}. } \label{f:mag} \end{figure} Figure~\ref{f:mag} shows the spatiotemporal image of the magnetization density $|\bm{F}|$ for the dynamics in Fig.~\ref{f:f_to_p1}. In the ferromagnetic region (red region in Fig.~\ref{f:mag}), the magnetization is $|\bm{F}| / \rho_0 \simeq 1$ except for that of the vortex cores (indicated by the arrows in Fig.~\ref{f:mag}). After the penetration of the phase boundary, the vortex cores in the $m = \pm 1$ components are occupied by the $m = \mp 1$ components, which results in the magnetization $F_z / \rho_0 \simeq \mp 1$ at the vortex cores. The rotation of the vortex dipoles in Fig.~\ref{f:f_to_p1}(g) is thus represented as double helices of magnetization in the spatiotemporal image in Fig.~\ref{f:mag}. The two helical structures in Fig.~\ref{f:mag} have opposite helicities. The results shown in Figs.~\ref{f:f_to_p1} and \ref{f:mag} can be understood as follows. A singly-quantized vortex in the ferromagnetic phase is written as \begin{equation} \label{ferrov} \left( \begin{array}{c} e^{\pm i\theta} \\ 0 \\ 0 \end{array} \right) f_1(r), \end{equation} where the origin of the polar coordinate is taken to be the center of the vortex core, and $f_1(r) \geq 0$ is a radial function with $f_1(0) = 0$ and $f_1(r)|_{r \gg \xi} = \rho_0^{1/2}$ with $\xi = \hbar / (M g_0 \rho_0)^{1/2}$. When the vortex in Eq.~(\ref{ferrov}) enters into the polar phase, it transforms to \begin{equation} \label{polarv} \rightarrow \frac{e^{\pm i\theta}}{\sqrt{2}} \left( \begin{array}{c} 1 \\ 0 \\ 1 \end{array} \right) f_{\rm p}(r) + \frac{1}{\sqrt{2}} \left( \begin{array}{c} 1 \\ 0 \\ -1 \end{array} \right) f_{\rm p}^{\rm core}(r), \end{equation} where the radial functions $f_{\rm p}(r), f_{\rm p}^{\rm core}(r) \geq 0$ satisfy $f_{\rm p}(0) = 0$, $f_{\rm p}(r)|_{r \gg \xi_s} = \rho_0^{1/2}$, and $f_{\rm p}^{\rm core}(r)|_{r \gg \xi_s} = 0$, in such a way that $f_{\rm p}^{\rm core}(r)$ occupies the core of $f_{\rm p}(r)$. The vortex states in Eqs.~(\ref{ferrov}) and (\ref{polarv}) can be smoothly connected by an intermediate state, \begin{eqnarray} \label{connect} & & e^{\pm i\theta} \left( \begin{array}{c} \cos\chi \\ 0 \\ \sin\chi \end{array} \right) {\cal F}[\chi; f_1(r) \rightarrow f_{\rm p}(r)] \nonumber \\ & & + \left( \begin{array}{c} -\sin\chi \\ 0 \\ \cos\chi \end{array} \right) {\cal F}[\chi; 0 \rightarrow f_{\rm p}^{\rm core}(r)], \end{eqnarray} where ${\cal F}$ smoothly connects the two functions as $\chi$ changes from 0 (ferromagnetic) to $\pi / 4$ (polar). The spin state of the core (the second terms in Eqs.~(\ref{polarv}) and (\ref{connect})) is taken to be such that the total density is independent of $\theta$. In fact, the total density is almost isotropic around the vortices, as shown in Figs.~\ref{f:f_to_p1}(a)--\ref{f:f_to_p1}(f). We consider the time evolution of Eq.~(\ref{polarv}). Since the first and second terms in Eq.~(\ref{polarv}) may have different energies, the time evolution of Eq.~(\ref{polarv}) is written as \begin{equation} \label{polarvt} \frac{e^{\pm i\theta}}{\sqrt{2}} \left( \begin{array}{c} 1 \\ 0 \\ 1 \end{array} \right) f_{\rm p}(r) + \frac{1}{\sqrt{2}} \left( \begin{array}{c} 1 \\ 0 \\ -1 \end{array} \right) f_{\rm p}^{\rm core}(r) e^{-i\delta\omega t}, \end{equation} where $\hbar \delta\omega$ is the energy difference between the core and the surrounding components. The magnetization of Eq.~(\ref{polarvt}) is calculated to be \begin{equation} \label{mag} |\bm{F}| = |F_z| = 2 f_{\rm p}(r) f_{\rm p}^{\rm core}(r) |\cos(\pm \theta + \delta \omega t)|, \end{equation} which has peaks at $r \simeq \xi_s$ and $\pm \theta + \delta\omega t = 0$ and $\pi$. Thus, a vortex that enters into the polar phase has two regions of nonzero magnetization around the core, and these rotate around one another at a frequency $\delta\omega$, resulting in the magnetization helices shown in Fig.~\ref{f:mag}. From Figs.~\ref{f:f_to_p1}(g) and \ref{f:mag}, $\delta\omega\tau / (2\pi)$ is found to be $\simeq -2$. This indicates that the energy of the core (the second term in Eq.~(\ref{polarv})) is smaller than that of the surrounding component (the first term in Eq.~(\ref{polarv})), since the core is not fully occupied, as shown in the total densities in Fig.~\ref{f:f_to_p1}, and hence the interaction energy is short of $g_0 \rho_0$ at the core. The rotating vortices shown in Fig.~\ref{f:f_to_p1}(g) can also be understood to be the result of the vortex-vortex interaction in an effectively two-component BEC. In a system with inhomogeneous density, a vortex with circulation $\bm{\kappa}$ moves in the direction of $\nabla \rho \times \bm{\kappa}$, i.e., in a direction perpendicular to the density gradient~\cite{AioiX}. Suppose that vortices in the $m = 1$ and $-1$ components are located at the origin and in its vicinity, respectively. The $m = \mp 1$ components occupy the vortex cores of $m = \pm 1$ components, and then the $m = -1$ density increases at the origin. Therefore, when a clockwise (counterclockwise) vortex in the $m = -1$ component exists around the origin, it experiences a density gradient toward the origin and moves around the origin clockwise (counterclockwise). Thus, clockwise (counterclockwise) vortices in the $m = 1$ and $-1$ components rotate around one another clockwise (counterclockwise), as observed in Fig.~\ref{f:f_to_p1}(g). A similar behavior can also be observed in a miscible two-component BEC (data not shown). \begin{figure}[tbp] \includegraphics[width=8cm]{fig3.eps} \caption{ (Color online) The spin states in Figs.~\ref{f:f_to_p1}(a)--\ref{f:f_to_p1}(f) are rotated by $\pi / 2$ about the $y$ axis. The upper and lower panels show the density profiles $|\psi_1|^2$ and $|\psi_0|^2$, respectively. The density $|\psi_{-1}|^2$ is similar to $|\psi_1|^2$. See the Supplemental Material for a movie of the dynamics of $|\psi_1|^2$. } \label{f:f_to_p2} \end{figure} It is interesting to see the behavior in Fig.~\ref{f:f_to_p1} on a different spin quantization axis. Figure~\ref{f:f_to_p2} shows the same dynamics as in Fig.~\ref{f:f_to_p1}, where $e^{-i f_y \pi / 2}$ is applied to the spin state (rotation by $\pi / 2$ about the $y$ axis). The ferromagnetic and polar states are rotated as $e^{-i f_y \pi / 2} (1, 0, 0)^T = (1/2, 1/\sqrt{2}, 1/2)^T$ and $e^{-i f_y \pi / 2} (1, 0, 1)^T / \sqrt{2} = (1, 0, 1)^T / \sqrt{2}$. After the vortex dipole penetrates through the phase boundary, it becomes a vortex dipole in the $m = \pm 1$ components, and the cores are occupied by the $m = 0$ component, as shown in Figs.~\ref{f:f_to_p1}(e) and \ref{f:f_to_p1}(f). Applying $e^{-i f_y \pi / 2}$ to Eq.~(\ref{polarvt}), we obtain \begin{equation} \label{rotate} \frac{e^{\pm i\theta}}{\sqrt{2}} \left( \begin{array}{c} 1 \\ 0 \\ 1 \end{array} \right) f_{\rm p}(r) + \left( \begin{array}{c} 0 \\ 1 \\ 0 \end{array} \right) f_{\rm p}^{\rm core}(r) e^{-i\delta\omega t}. \end{equation} Unlike the behavior shown in Fig.~\ref{f:f_to_p1}(g), the rotation dynamics of vortex dipoles is not observed with this spin quantization axis. The magnetization of Eq.~(\ref{rotate}) is of course the same as in Eq.~(\ref{mag}). \begin{figure}[tbp] \includegraphics[width=8cm]{fig4.eps} \caption{ (Color online) Time evolution of a vortex dipole traveling from the ferromagnetic phase to the polar phase, where the initial distance between the vortices is $\Delta = 3.2\xi_s$. The upper and lower panels show the density profiles $|\psi_1|^2$ and $|\psi_{-1}|^2$, respectively. The density $|\psi_0|^2$ is always negligible. The arrows indicate the directions in which the vortices circulate. The window of each panel is $-5 \xi < x < 5 \xi$ and $-10 \xi < y < 10 \xi$. See the Supplemental Material for a movie of the dynamics of $|\psi_1|^2$. } \label{f:hqv} \end{figure} Figure~\ref{f:hqv} shows the generation of a half-quantum vortex dipole, where the distance in the initial vortex dipole $\Delta = 3.2 \xi_s$ is larger than that in Figs.~\ref{f:f_to_p1}--\ref{f:f_to_p2}. As they pass through the phase boundary, the size of the vortex cores with the $m = 1$ components significantly expand and are occupied by the $m = -1$ components, yielding the half-quantum vortex dipole in the polar phase. The transformation from a singly-quantized vortex in the ferromagnetic phase to a half-quantum vortex in the polar phase is expressed as \begin{eqnarray} \label{hqv1} \left( \begin{array}{c} e^{\pm i\theta} \\ 0 \\ 0 \end{array} \right) f_1(r) & \rightarrow & \left( \begin{array}{c} e^{\pm i\theta} \cos\chi \\ 0 \\ \sin\chi \end{array} \right) {\cal F}[\chi; f_1(r) \rightarrow f_{\rm hqv}(r)] \nonumber \\ & & + \left( \begin{array}{c} 0 \\ 0 \\ 1 \end{array} \right) {\cal F}[\chi; 0 \rightarrow f_{\rm hqv}^{\rm core}(r)] \nonumber \\ & \rightarrow & \frac{1}{\sqrt{2}} \left( \begin{array}{c} e^{\pm i\theta} \\ 0 \\ 1 \end{array} \right) f_{\rm hqv}(r) + \left( \begin{array}{c} 0 \\ 0 \\ 1 \end{array} \right) f_{\rm hqv}^{\rm core}(r), \nonumber \\ \end{eqnarray} where the radial functions $f_{\rm hqv}(r), f_{\rm hqv}^{\rm core}(r) \geq 0$ satisfy $f_{\rm hqv}(0) = 0$, $f_{\rm hqv}(r)|_{r \gg \xi_s} = \rho_0^{1/2}$, and $f_{\rm hqv}^{\rm core}(r)|_{r \gg \xi_s} = 0$, i.e., $f_{\rm hqv}^{\rm core}(r)$ occupies the core of $f_{\rm hqv}(r)$. Thus, when a vortex dipole created in the ferromagnetic phase enters into the polar phase, it transforms into one of two kinds of spin-vortex dipoles, as shown in Figs.~\ref{f:f_to_p1} and \ref{f:hqv}, depending on the distance $\Delta$ in the initial vortex dipole. The leapfrogging ferromagnetic-core vortex dipoles are generated for $\Delta \lesssim 2 \xi_s$, and the half-quantum vortex dipole is generated for $\Delta \gtrsim 2 \xi_s$. From a topological point of view, in general, vortex states in the ferromagnetic phase, classified by $\pi_1(G_f) = \mathbb{Z}_2$, and those in the polar phase, classified by $\pi_1(G_p) = \mathbb{Z}$, can be smoothly connected to each other~\cite{Kobayashi}. In fact, spin states far from the core may be transformed from the ferromagnetic phase to the polar phase as \begin{equation} \label{general} \left( \begin{array}{c} e^{\pm i\theta} \\ 0 \\ 0 \end{array} \right) \rightarrow \left( \begin{array}{c} e^{\pm i\theta} \cos\chi \\ 0 \\ e^{i n \theta} \sin\chi \end{array} \right) \rightarrow \frac{1}{\sqrt{2}} \left( \begin{array}{c} e^{\pm i\theta} \\ 0 \\ e^{i n \theta} \end{array} \right), \end{equation} where $\chi$ changes from 0 to $\pi / 4$, and $n$ is an integer. The dynamics shown in Figs.~\ref{f:f_to_p1} and \ref{f:hqv} correspond, respectively, to $n = \pm 1$ and $n = 0$ in Eq.~(\ref{general}). The transformation with other values of $n$ are not realized in the present dynamics. \section{Propagation of a vortex dipole from polar to ferromagnetic phase} \label{s:p_to_f} \begin{figure}[tbp] \includegraphics[width=8cm]{fig5.eps} \caption{ (Color online) Time evolution of a vortex dipole traveling from the polar phase to the ferromagnetic phase, where the initial distance between the vortices is $\Delta = 1.6\xi_s$. (a)--(f) show $|\psi_1|^2$, $|\psi_{-1}|^2$, and $\sum_m |\psi_m|^2$, where $|\psi_0|^2$ is always negligible. The arrows in (a) indicate the directions in which the vortices circulate. (g) Time evolution of the density profiles $|\psi_1|^2$ and $|\psi_0|^2$, where the spin quantization axis is rotated by $\pi / 2$ about the $y$ axis. The windows of each panel are (a)--(f) $-4 \xi_s < x < 4 \xi_s$ and $-3 \xi_s < y < 3 \xi_s$, and (g) $-4.5 \xi_s < x < -1.5 \xi_s$ and $-3 \xi_s < y < 3 \xi_s$. See the Supplemental Material for a movie of the dynamics of $|\psi_{-1}|^2$ in (a)--(f) and that of $|\psi_1|^2$ in (g). } \label{f:p_to_f} \end{figure} We next consider cases in which a vortex dipole is created in the polar phase ($x > 0$) and moves toward the ferromagnetic phase ($x < 0$). In Fig.~\ref{f:p_to_f}, a vortex dipole is created at $z_\pm = 5 \xi_s \mp i \Delta / 2$ with $\Delta = 1.6\xi_s$, which then moves in the $-x$ direction. When the vortex dipole passes through the phase boundary, a complicated spin texture is formed, as shown in Figs.~\ref{f:p_to_f}(c) and \ref{f:p_to_f}(d). After that, on the ferromagnetic side, a vortex dipole is transferred into the $m = 1$ component, whose cores are occupied by the $m = -1$ component, as shown in Fig.~\ref{f:p_to_f}(f). Unlike the case in Fig.~\ref{f:f_to_p1}(f), the core size of the vortex dipole traveling in the ferromagnetic phase is unchanged. Figure~\ref{f:p_to_f}(g) shows the dynamics after passing through the phase boundary, where the spin quantization axis is rotated by $\pi / 2$ about the $y$ axis ($e^{-i f_y \pi / 2}$ is applied). In this quantization axis, the vortex dipole in Fig.~\ref{f:p_to_f}(f) behaves as leapfrogging vortex dipoles in the $m = \pm 1$ and 0 components. The vortex transformation from that in Fig.~\ref{f:f_to_p1}(a) to that in Fig.~\ref{f:f_to_p1}(f) is expressed as \begin{eqnarray} \label{pf} \frac{e^{\pm i\theta}}{\sqrt{2}} \left( \begin{array}{c} 1 \\ 0 \\ 1 \end{array} \right) f_1(r) & \rightarrow & e^{\pm i\theta} \left( \begin{array}{c} \cos\chi \\ 0 \\ \sin\chi \end{array} \right) {\cal F}[\chi; f_1(r) \rightarrow f_{\rm f}(r)] \nonumber \\ & & + \left( \begin{array}{c} -\sin\chi \\ 0 \\ \cos\chi \end{array} \right) {\cal F}[\chi; 0 \rightarrow f_{\rm f}^{\rm core}(r)] \nonumber \\ & \rightarrow & e^{\pm i\theta} \left( \begin{array}{c} 1 \\ 0 \\ 0 \end{array} \right) f_{\rm f}(r) + \left( \begin{array}{c} 0 \\ 0 \\ 1 \end{array} \right) f_{\rm f}^{\rm core}(r), \nonumber \\ \end{eqnarray} where $\chi$ changes from $\pi/4$ to 0. The radial functions $f_{\rm f}(r), f_{\rm f}^{\rm core}(r) \geq 0$ satisfy $f_{\rm f}(0) = 0$, $f_{\rm f}(r)|_{r \gg \xi_s} = \rho_0^{1/2}$, and $f_{\rm f}^{\rm core}(r)|_{r \gg \xi_s} = 0$. In the final state in Eq.~(\ref{pf}), which corresponds to Fig.~\ref{f:p_to_f}(f), the spin state has the form of $\propto (e^{\pm i\theta}, 0, 1)^T$ for a radius $r \sim \xi_s$ satisfying $f_{\rm f}(r) \simeq f_{\rm f}^{\rm core}(r)$, which is a half-quantum vortex~\cite{Kobayashi}. Thus, the vortex dipole in the ferromagnetic phase shown in Fig.~\ref{f:p_to_f}(f) contains a half-quantum vortex dipole around the ferromagnetic cores. Applying $e^{-i f_y \pi / 2}$ to the final state in Eq.~(\ref{pf}) and multiplying the second term by the factor $e^{-i \delta \omega t}$, we obtain \begin{eqnarray} \label{ferrovt} & & e^{\pm i\theta} \left( \begin{array}{c} 1 / 2 \\ 1 / \sqrt{2} \\ 1 / 2 \end{array} \right) f_{\rm f}(r) + \left( \begin{array}{c} 1 / 2 \\ -1 / \sqrt{2} \\ 1 / 2 \end{array} \right) f_{\rm f}^{\rm core}(r) e^{-i\delta\omega t} \nonumber \\ & = & e^{-i\delta\omega t} \left( \begin{array}{c} \left[ e^{i(\pm\theta + \delta\omega t)} f_{\rm f}(r) + f_{\rm f}^{\rm core}(r) \right] / 2 \\ \left[ e^{i(\pm\theta + \delta\omega t)} f_{\rm f}(r) - f_{\rm f}^{\rm core}(r) \right] / \sqrt{2} \\ \left[ e^{i(\pm\theta + \delta\omega t)} f_{\rm f}(r) + f_{\rm f}^{\rm core}(r) \right] / 2 \end{array} \right). \end{eqnarray} Equation~(\ref{ferrovt}) has vortex cores at $r \sim \xi_s$ and $\pm\theta + \delta\omega t = \pi$ in the $m = \pm 1$ component, and at $\pm\theta + \delta\omega t = 0$ in the $m = 0$ component. Thus, each component of Eq.~(\ref{ferrovt}) has a vortex core, and the vortices in the $m = \pm 1$ and $m = 0$ components rotate around one another with frequency $\delta\omega$, which explains the behavior in Fig.~\ref{f:p_to_f}(g). Since the positively (negatively) charged vortices rotate around one another counterclockwise (clockwise) in Fig.~\ref{f:p_to_f}(g), $\delta\omega$ is found to be negative, in a manner similar to the case of Figs.~\ref{f:f_to_p1} and \ref{f:mag}. \begin{figure}[tbp] \includegraphics[width=8cm]{fig6.eps} \caption{ (Color online) Time evolution of a system in which a half-quantum vortex dipole is imprinted to the polar side of the initial state. In (a), singly-quantized vortices are imprinted only to the $m = 1$ wave function, and in (b), they are imprinted only to the $m = -1$ wave function. The distances between the vortices in the initial vortex dipole are (a) $\Delta = 6.4\xi_s$ and (b) $\Delta = 3.2\xi_s$. The arrows indicate the directions in which the vortices circulate. The window of each panel is $-5 \xi_s < x < 5 \xi_s$ and $-10 \xi_s < y < 10 \xi_s$. See the Supplemental Material for movies of the dynamics of $|\psi_1|^2$ in (a) and that of $|\psi_{-1}|^2$ in (b). } \label{f:hqv2} \end{figure} We examine the dynamics of a half-quantum vortex dipole that is created on the polar side and moves toward the ferromagnetic side. In Fig.~\ref{f:hqv2}(a), vortices are imprinted only to the $m = 1$ component, using Eq.~(\ref{phase}) with $z_{\pm} = 5\xi \mp 3.2 \xi_s$. As the half-quantum vortex dipole moves in the $-x$ direction and approaches the phase boundary, the distance between the vortex and antivortex is reduced ($t = 21 \tau$ in Fig.~\ref{f:hqv2}(a)). On the ferromagnetic side, the vortex and antivortex merge into a single density hole occupied by the $m = -1$ component, which continues to move in the $-x$ direction ($t = 26 \tau$ in Fig.~\ref{f:hqv2}(a)). This is quite different from Fig.~\ref{f:p_to_f}(f), where the vortex and antivortex survive on the ferromagnetic side. The different behaviors originate from the different energy scales of the vortices. It follows from the core sizes that the energies of the singly quantized vortices in Fig.~\ref{f:p_to_f}(a) are larger than those of the spin vortices in Fig.~\ref{f:p_to_f}(f), whereas the energies of the half-quantum vortices in Fig.~\ref{f:hqv2}(a) are smaller. Thus, the half-quantum vortex dipole cannot transform into a spin vortex dipole on the ferromagnetic side due to a lack of energy. Figure~\ref{f:hqv2}(b) shows the case in which the phase is imprinted only to the $m = -1$ initial state, using Eq.~(\ref{phase}) with $z_{\pm} = 5\xi \mp 1.6 \xi_s$. In this case, the penetration of a vortex dipole is topologically prohibited, since the vortices in the $m = -1$ component cannot be smoothly connected to those in the $m = 1$ component. The half-quantum vortex dipole therefore does not penetrate the phase boundary, but instead the vortex and antivortex disintegrate and move along the phase boundary in opposite directions, as shown in Fig.~\ref{f:hqv2}(b). \section{Conclusions} \label{s:conc} We have investigated the dynamics of vortex dipoles across the boundary between the ferromagnetic and polar phases in a spin-1 BEC. We numerically solved the Gross--Pitaevskii equation and found a rich variety of spin dynamics. When a singly-quantized vortex dipole is created on the ferromagnetic side and propagated toward the polar side, it transforms into one of two kinds of spin-vortex dipoles, depending on its velocity. For a large velocity, ferromagnetic-core vortex dipoles are created in the $m = 1$ and $-1$ components, which exhibit the leapfrogging dynamics, as shown in Figs.~\ref{f:f_to_p1} and \ref{f:mag}. For a small velocity, a half-quantum vortex dipole is generated, as shown in Fig.~\ref{f:hqv}. When a singly-quantized vortex dipole created on the polar side moves into the ferromagnetic side, it transforms into a vortex dipole whose ferromagnetic cores are surrounded by half-quantum vortices, as shown in Figs.~\ref{f:p_to_f}(a)--\ref{f:p_to_f}(f). This state also exhibits the leapfrogging behavior on a different quantization axis, as shown in Fig.~\ref{f:p_to_f}(g). When a half-quantum vortex dipole is created on the polar side and propagated to the ferromagnetic side, it coalesces into a low energy droplet or is turned away at the phase boundary, as shown in Fig.~\ref{f:hqv2}. We have considered an ideal system that is infinite and homogeneous. In a realistic experimental system confined in a trapping potential, the system size must be much larger than a vortex dipole and its trajectory. An elongated oblate BEC, as used in the Berkeley experiment~\cite{Sadler}, would be suitable. An initial vortex dipole can be created by an external laser beam shifting in a BEC~\cite{Neely,AioiX}. A half-quantum vortex-dipole, as shown in Fig.~\ref{f:hqv2}, may be created by an $m$-dependent external potential~\cite{Ji,Chiba}. Our study presents an example of the dynamical transformation of topological defects, in which topological structures are dynamically changed. It will also be interesting to consider collisions of topological defects. During collisions, topological structures may be dynamically changed, and different topological structures may be scattered after the collisions, which might be relevant to the collisions of elementary particles. \begin{acknowledgments} This work was supported by JSPS KAKENHI Grant Number 26400414 and by KAKENHI (No. 25103007, ``Fluctuation \& Structure'') from MEXT, Japan. \end{acknowledgments}
\section{Introduction} Distributed decision-making in robotic networks is a ubiquitous problem, with applications as diverse as state estimation \cite{ROS:07}, formation control \cite{WR-RWB-EMA:07}, tracking \cite{YH-JH-LG:06} and cooperative task allocation \cite{MdW-BC:09}. Distributed decision-making problems such as leader election, majority voting, distributed hypothesis testing, and some distributed optimization problems can all be abstracted as instances of the \emph{consensus} problem, where nodes in a robotic network have to agree on some common value \cite{JNT-MA:85,NL:96}. For this reason, the consensus problem has gathered significant interest in the Control Systems community in recent years, following the seminal works in \cite{JNT:84-extra, ROS-RMM:03c,AJ-JL-ASM:02}. Research in the Control Systems community has mainly focused on the \emph{average} consensus problem and, in particular, on local averaging algorithms, where nodes repeatedly average their state with their neighbors': fundamental limitations on the \emph{time} complexity of local averaging algorithms are now known \cite{AO-JNT:07}. Conversely, research in the Computer Science community has mainly focused on lower bounds on the general consensus problem in presence of node failures (both non-malicious and byzantine): several lower bounds on the time and communication complexity of the general consensus problem are now known. Significant attention has also been devoted to the \emph{leader election} problem, a specific consensus problem where agents in a network select a single agent as their leader. Fundamental limitations of the leader election problem in terms of time and communication complexity and matching optimal algorithms are now well-understood \cite{NL:96}. However, complexity results in the Control Systems and in the Computer Science communities strongly rely on the assumption that all messages can be delivered within a \emph{finite} amount of time that does \emph{not} depend on message size \cite{NL:96}. This assumption has significant effects on the \emph{frequency bandwidth} collectively required by the agents: however, despite the large interest in the consensus problem and its applications, the problem of bandwidth use has seen very limited investigation in both the Computer Science and the Control Systems communities. \emph{Motivation}: In this paper we argue that bandwidth use can play a significant role in the real-world performance of consensus algorithms and significantly limit their scalability as the number of agents increases. In order to appreciate the importance of bandwidth use in modern robotic networks, consider the following scenario. A network of $n=100$ robotic agents, each with six mechanical degrees of freedom is tasked with averaging their state (i.e. performing average consensus), e.g. to control their formation \cite{WR-RWB-EMA:07} or to filter noisy observations of a common target \cite{ROS:07}. Each agent's state or observation can be represented by twelve floating point numbers: the size of each agent's initial condition is $b=768$ bits. Each message should be delivered in $10$ ms at most, in order to guarantee acceptable time performance. The complexity results in Section \ref{sec:algs} allow us to show that a bandwidth-optimal algorithm such as GHS with convergecast achieves consensus with a bandwidth complexity of 143 kbps (and converges in approximately 6.7 s). The popular average-based consensus algorithm requires 7750 kbps and has a significantly slower convergence rate, achieving convergence in approximately 100 s. Finally, the time-optimal flooding algorithm only requires 1 s to converge, but it requires a bandwidth of 775 Mbps. As a comparison, a bandwidth-optimized protocol such as 802.11n WiFi requires use of the entire 2.4 GHz ISM band to transmit 288Mbps over very short distances: thus, selecting an inappropriate consensus algorithm can have a dramatic impact on the real-world performance, the scalability and potentially even the stability of a cyber-physical network. \emph{Statement of contributions}: The contribution of this paper is threefold. First, we propose a rigorous metric for the bandwidth complexity of decentralized algorithms with omnidirectional (broadcast) communication channels and discuss its relevance to modern media access control (MAC) mechanisms. Second, we prove a lower bound on the bandwidth complexity of the generalized consensus problem. The bound is tight for a wide class of consensus functions that includes many distributed consensus problems such as mean and weighed mean, leader election, selected distributed optimization problems and majority voting. Finally, we show that the hybrid algorithm proposed by the authors in \cite{FR-MP:13} achieves intermediate bandwidth performance between the time-optimal flooding algorithm and the bandwidth-optimal GHS algorithm with convergecast. This result, combined with our previous findings, shows that our hybrid algorithm can be tuned to satisfy mixed time-bandwidth metrics as well as mixed time-energy metrics, trading time complexity (and, to an extent, robustness) for spectrum utilization and energy consumption. A graphical depiction of our lower bounds on time, byte and bandwidth complexity of the flooding algorithm, the GHS algorithm with convergecast and our hybrid algorithm is shown in Figure \ref{fig:bounds}. \begin{figure*}[h!tp] \centering \begin{subfigure}[h]{0.3\textwidth} \includegraphics[width=\textwidth]{Time_bound.pdf} \caption{Bounds on time complexity} \end{subfigure} \begin{subfigure}[h]{0.3\textwidth} \includegraphics[width=\textwidth]{Byte_bound.pdf} \caption{Bounds on byte complexity} \end{subfigure} \begin{subfigure}[h]{0.3\textwidth} \includegraphics[width=\textwidth]{Bandwidth_bound.pdf} \caption{Bounds on bandwidth complexity} \end{subfigure} \caption{Bounds on time, byte and bandwidth complexity of the flooding algorithm (in blue, dashed), GHS algorithm with convergecast (in green, dash-dotted) and of our hybrid algorithm (in red, solid) as a function of the tuning parameter $m$ for a network of size $n=100$ with state size $b=768$ bits. Our hybrid algorithm recovers the time-optimal performance of the flooding algorithm as $m\to n$. Conversely, the hybrid algorithm recovers the byte and bandwidth performance of the byte-optimal and bandwidth-optimal GHS algorithm with convergecast, presented in Section \ref{sec:algs}, as $m\to 1$.} \label{fig:bounds} \end{figure*} \emph{Organization}: This paper is organized as follows. In Section \ref{sec:setup}, after formalizing our agent model and network model, we give a formal definition of bandwidth complexity and justify its relevance to modern multi-agent media access control mechanisms. We then provide a rigorous definition of the generalized consensus problem. In Section \ref{sec:lbs} we prove a lower bound on the bandwidth complexity of the generalized consensus problem. In Section \ref{sec:algs} we prove tightness of the lower bound. We also study the bandwidth complexity of common consensus algorithms (including the flooding algorithm, the GHS algorithm with convergecast and the average-based consensus-algorithm). A variation of the GHS algorithm with convergecast achieves the lower bound presented in the previous section. In Section \ref{sec:tunable}, we study the bandwidth complexity of the hybrid algorithm introduced by the authors in \cite{FR-MP:13}. We show that the bandwidth complexity of the algorithm smoothly transitions from bandwidth-optimal performance to the performance of the time-optimal flooding algorithm. Finally, in Section \ref{sec:conclusions}, we draw our conclusions and discuss directions for future research. \section{Problem setup} \label{sec:setup} In this section we introduce the model and the complexity metrics used in this work. We also describe naming conventions that will be used in the rest of this paper. A preliminary version of the model has appeared in \cite{FR-MP:13}. \subsection{Agent model} An agent in a robotic network is modeled as an input/output (I/O) automaton, i.e., a labeled state transition system able to send messages, react to received messages and perform arbitrary internal transitions based on the current state and messages received. A precise definition of I/O automaton is provided in \cite[pp. 200-204]{NL:96} and is omitted here in the interest of brevity. All agents in a robotic network are identical except for a unique identifier (UID - for example, an integer). The time evolution of each agent is characterized by two key assumptions: \begin{itemize} \item {\bf Fairness assumption}: the order in which transitions happen and messages are delivered is not fixed a priori. However, any enabled transition will \emph{eventually} happen and any sent message will \emph{eventually} be delivered. \item {\bf Non-blocking assumption}: every transition is activated within $l$ time units of being enabled and every message is delivered within $d$ time units of being dispatched. \end{itemize} Essentially, the fairness assumption states that each agent will eventually have an opportunity to perform transitions, while the non-blocking assumption gives timing guarantees (but no synchronization). We refer the interested reader to \cite[pages 212-215]{NL:96} for a detailed discussion of these assumptions. We argue here that these are \emph{minimal} assumptions for most real-world robotic networks. \subsection{Network model} A \emph{robotic network} comprising $n$ agents is modeled as a \emph{connected}, \emph{undirected} graph $G = (V,E)$, where $V = \{1,\ldots, n\}$ is the node set, and $E\subset V\times V$, the edge set, is a set of \emph{unordered} node pairs modeling the availability of a communication channel. Two nodes $i$ and $j$ are neighbors if $(i, j)\in E$. The neighborhood set of node $i\in V$ is the set of nodes $j\in V$ that are neighbors of node $i$. Henceforth, we will refer to nodes and agents interchangeably. Our model is \emph{asynchronous}, i.e., computation steps within each node and communication are, in general, asynchronous. In this paper we focus on \emph{static networks}, i.e., robotic networks where the edge set does not change during the execution of an algorithm. However, we do remark that (i) our lower bounds also apply to time-varying networks (although, of course, they may not be tight) and (ii) the hybrid algorithm described in section \ref{sec:tunable} has \emph{limited, tunable} resistance to network disruption, requiring a tunable amount of time to reconfigure after network disruptions. We refer the interested reader to \cite{FR-MP:13} for a thorough discussion of the recovery mechanism and of its time complexity. \subsection{Model of communication} Nodes communicate with their neighbors according to a \emph{local broadcast} communication scheme. A node can send a message to all neighbors simultaneously: the cost of a message (in terms of energy consumption and bandwidth use) is independent of the number of receivers. We remark that local broadcast algorithms can emulate one-to-one communication: it is sufficient to append the intended recipient's UID to each broadcast and instruct non-recipients to ignore the message. The local broadcast communication scheme is representative of robotic networks where agents are equipped with \emph{omnidirectional} antennas: this arrangement is typical of most current airborne and ground-based robotic networks, where steerable antennas are unadvisable due to the agents' mobility. Message transmission requires a finite, nonvanishing amount of time; the non-blocking assumption ensures that every message is delivered within time $d$. We consider $d$ to be constant: that is, all messages are delivered within a maximum time that does \emph{not} depend on message size or type. This assumption is widely used in the Computer Science community \cite{NL:96} and is typical of TCP-like communication protocols. Thus, the parameter $d$ represents a \emph{desired performance level}. \emph{Collisions} occur when (part of) two or more messages are transmitted on the same frequency at the same time: when a collision occurs, all messages involved in the collision are not delivered. When analyzing bandwidth complexity, we assume that messages are sent at a \emph{constant rate} throughout a window of length $d$: it is easy to see that ``bursty'' transmission would only decrease bandwidth performance. We remark that, in presence of a TCP communication protocol, collisions do not cause messages to be permanently lost: nodes can sense the collision and resend the information at a later time. However, frequent collisions can have a major impact on the time required to deliver a message and, as a result, on the execution time of an algorithm. We also remark that, in \emph{directional} communication schemes, communication channels can be \emph{spatially} separated: that is, two messages may be transmitted on the same frequency at the same time with minor interference if the respective directional communication channels do not physically overlap. In our omnidirectional communication scheme, on the other hand, collisions occur whenever two nodes within range of one another send messages on the same frequency at the same time; in addition, even if two nodes are not within range of one another, collisions may occur if they are trying to contact a third node in range of both (this problem is known as the ``hidden node problem'' in the telecommunication community). In this paper we assume no restrictions to the network topology the nodes can assume: we remark that there exist both dense and sparse network topologies where every node's communications may spatially interfere with all other nodes'. \subsection{Bandwidth complexity measure} We define bandwidth complexity as the infimum worst-case (over graph topologies, initial values, fair executions and execution time) overall number of bytes transmitted at the same instant by all agents in the network. Let $\mathcal G$ be a set of graphs with node set $V=\{1,\ldots, n\}$. For a given graph $G\in \mathcal G$, let $\mathcal F(a, x, G)$ be the set of \emph{fair executions} for an algorithm $a\in \mathcal A$ and a set of initial conditions $x \in \mathcal X^{n}$ (a fair execution is an execution of an algorithm that satisfies the fairness and non-blocking assumptions stated above). Rigorously, the bandwidth complexity for a given consensus function $f$ with respect to the class of graphs $\mathcal G$ is \begin{flalign*} &\textrm{FC}(f, G):=&\\ & \inf_{a\in \mathcal A} \, \sup_{G\in \mathcal G}\, \sup_{x \in \mathcal X^{|G|}} \, \sup_{\alpha \in \mathcal F(a, x,G)} \, \sup_{t\in [0, T(a, x, \alpha, G)]} F(a, x, \alpha, G,t) \end{flalign*} where $F(a, x, \alpha, G,t)$ is the bandwidth (measured by the size of all messages transmitted at time $t$ divided by the maximum transmission time $d$) at time $t$ of execution $\alpha$ of algorithm $a$ with initial conditions $x$ on a graph $G$. While very simple, the bandwidth complexity measure is a reliable proxy for many wireless communication protocols and media access control (MAC) mechanisms. Its interpretation varies depending on the specific MAC mechanism employed: \begin{itemize} \item If Frequency Division Multiple Access (FDMA) is employed, the frequency bandwidth required by a single message is proportional to the message size divided by the maximum transmission time $d$; the overall frequency bandwidth required by the network is proportional to the maximum \emph{overall} size of messages being transmitted at a given time divided by the maximum transmission time $d$, since every message must be broadcast on a different frequency slot; \item If a Time Division Multiple Access (TDMA) media access control mechanism is employed, each agent is allocated a time slot in a round robin fashion so that only one agent can transmit during a given time slot. The sum of the durations of all time slots must be smaller than $d$ to guarantee that all sent messages be delivered within $d$ time units: thus, in order to guarantee a timely delivery, the frequency bandwidth required must be proportional to the maximum overall size of all messages sent at any instant of time, divided by $d$. \item If Code Division Multiple Access (CDMA) is employed, multiple messages are relayed on the same, wide frequency spectrum at the same time; a \emph{spread spectrum} technique is employed to make decoding of sent messages possible. The bandwidth of the spread spectrum is significantly larger than the bandwidth of the uncoded signal: in particular (i) the bandwidth required by a single message before encoding is proportional to its size (in bytes) and (ii) the spreading gain is roughly proportional to the maximum number of users that the network can support. Thus, if all agents transmit messages of the same size at the same time (as will be the case for the proof of the lower bound we study in this paper and also for the algorithms we discuss in Section \ref{sec:algs}), bandwidth complexity captures the frequency spectrum required for successful communication with a CDMA MAC mechanism. \item A rigorous study of the effect of available bandwidth on channel capacity when collision-detection mechanisms such as CSMA/CA are employed is beyond the scope of our work. However we remark that, for a given message size, increasing bandwidth reduces the time required to transmit a message and thus the network load, significantly reducing the probability of collisions and thus increasing the effective throughput. \end{itemize} Furthermore, regardless of the MAC mechanism employed, for large signal-to-noise ratios, the maximum capacity of a wireless channel is approximately proportional to bandwidth, as shown by Shannon in \cite{CES:49}. \subsection{Model of computation} In this paper we study collective decision-making problems where each node in the network is endowed with an initial value $x_i$ (which can be represented with $b$ bits) and \emph{each} node should output the value of a function of the initial values of \emph{all} nodes. In other words, each agent, after exchanging messages with its neighbors and performing internal state transitions, should output $f(x_1, \ldots, x_n)$ for some computable function $f$, which we call a \emph{consensus} function. We formalize the notions of consensus function as follows. \textbf{Consensus functions}: A consensus function is a \emph{computable} function $f : \mathcal X^n \mapsto \real$ that depends on \emph{all} its arguments. More precisely, for each element $x = (x_1, \ldots, x_n) \in \mathcal X^n$ and for all $i\in\{1, \ldots, n\}$ one can find elements $x_{i}^{(1)}\in \mathcal X$ and $x_{i}^{(2)}\in \mathcal X$ such that \[ f(x_1, \ldots, x_{i}^{(1)}, \ldots, x_n)\neq f(x_1, \ldots, x_{i}^{(2)}, \ldots, x_n). \] Loosely speaking, such choice of consensus function implies that each node is needed for the decision-making process. We collectively refer to problems involving the distributed computation of consensus functions (as defined above) as \emph{generalized consensus}. We introduce a \emph{representation} property for consensus functions that will be instrumental to derive fundamental limitations of performance in terms of the amount of information exchanged. \textbf{Hierarchically computable consensus function}: A consensus function is hierarchically computable if it can be written as the composition of a commutative and associative binary operator $\ast$, that is \[ f(x_1, x_2,\ldots, x_n) = x_1 \ast x_2\ast \ldots\ast x_n. \] (The name is inspired by the observation that hierarchically computable functions can be computed with messages of small size on a \emph{hierarchical} structure such as a tree). Furthermore, for a consensus function to be hierarchically computable, the consensus value as well as all intermediate products should be of the same size (in bytes) as the initial value. That is, if storing $x_i$ requires $\Theta(b)$ bytes, then storing the result of the operation $(x_i \ast x_j)$ and of the consensus value $f(x_1, \ldots, x_n)$ should also require $\Theta(b)$ bytes. We remark that the class of hierarchically computable consensus functions includes average and weighed average, MAX and MIN (often used in leader election), voting and selected distributed optimization problems. We refer the reader to \cite{FR-MP:14a} for a more exhaustive characterization of this class of functions. \subsection{Nomenclature} In the rest of this paper, we use the following definitions for nodes belonging to a rooted tree structure. \begin{itemize} \item Each rooted tree contains a \emph{root node}. \item A node $j$ is the \emph{child} of node $i$ if (i) the shortest path from node $j$ to the root is one edge longer than the shortest path from node $i$ to the root and (ii) node $i$ and node $j$ share an edge in the tree. Conversely, node $i$ is called node $j$'s \emph{parent}. \item A \emph{leaf node} is a node with no children. \item A node $j$ is a \emph{descendant} of node $i$ if (i) the shortest path from node $j$ to the root is strictly longer than the distance from node $i$ to the root and (ii) the path connecting node $j$ and the root node contains node $i$. \item The set containing a node $j$ and all its descendants is the \emph{branch} of node $j$. \end{itemize} Figure \ref{fig:treenomenclature} shows a graphical depiction of the definitions above. \begin{figure}[h] \centering \includegraphics[width=.45\textwidth]{Tree_nomenclature.pdf} \caption{Naming conventions for nodes belonging to a rooted tree structure} \label{fig:treenomenclature} \end{figure} \section{Lower bounds on bandwidth complexity of consensus} \label{sec:lbs} In this section, we present two lower bounds on the bandwidth complexity of the consensus problem. The first bound applies to \emph{asynchronous} executions and only depends on very mild assumptions on the minimum number of available UIDs; the second bound applies to synchronous as well as asynchronous executions and requires slightly more restrictive assumptions on the execution time and on the minimum number of available UIDs. \begin{proposition}[Lower bound on the bandwidth complexity of the consensus problem in asynchronous executions] \label{prop:asynclb} Assume that messages carry a sender and/or a receiver ID. Assume also that agents' UIDs are selected from a set $S$ of cardinality $|S|\geq 2n$. Then, for a given consensus function $f$ and class of graphs $\mathcal G$ with $n$ nodes, $\textrm{FC}(f, \mathcal G) \in \Omega((n\log n +b)/d)$. \end{proposition} \begin{proof} When they start execution, agents decide whether to send a message or whether to wait in silence until they hear a message. Since agents have no information about other nodes before they receive a message, their decision is based solely only on their UID and, possibly, their initial condition. We consider asynchronous executions: agents do not have access to a shared clock and therefore can not make decisions based on time. We wish to show that, if the agents' UIDs are drawn from a set $U$ with cardinality $U\geq 2n$, then there exist a subset $M\subseteq U$ with cardinality $|M|\geq n$ such that all agents with an UID from $M$ send a message before receiving a message at least for one set of initial conditions. We proceed by contradiction. Call $S=U\setminus M$ the set of UIDs such that, for all initial conditions, an agent sends no message before receiving one message. Now, assume by contradiction that the cardinality of $M$ is smaller than $n$. Then the cardinality of $S$ is larger than $n$: there exists an assignment of $n$ UIDs to the agents such that no node in the network sends a message before receiving one. Any network formed from these agents exchanges no messages and, in particular, it fails to solve the consensus problem unless all initial conditions are identical. We have reached a contradiction. If the cardinality of $M$ is larger than $n$, then an adversary can select $n$ UIDs and $n$ matching initial conditions from $M$ and form a network where every node sends a message before receiving one. In an asynchronous executions, all nodes can transmit their first message simultaneously. Furthermore, at least $\log n$ bits are required to store $n$ distinct UIDs: thus, if messages contains the transmitter or the receiver UID, the size of each message is lower-bounded by $\log n$. Therefore $n\log n$ bits may be transmitted simultaneously, with a bandwidth complexity of $n\log n/d$. Finally, we observe that every agent transmits its initial value at least once. If this were not the case, then at least one agent would never inform other nodes of its initial value: then, since the consensus function is sensitive to all initial values, the algorithm would be unable to correctly compute the consensus function. Transmission of an initial value requires $(\log n+b)$ bytes: its bandwidth complexity is $(\log n + b)/d$. We can then conclude that the bandwidth complexity of the consensus problem is $\textrm{FC}(f, \mathcal G) \in \Omega((n\log n +b)/d)$. \end{proof} Proposition \ref{prop:asynclb} strongly relies on the assumption of asynchronous communication. In the next proposition we show that the lower bound on bandwidth complexity also applies to synchronous executions if (i) the pool of UIDs is ``large enough'' and (ii) the algorithm is required to terminate in a bounded number of steps. \begin{proposition}[Lower bound on the bandwidth complexity of the consensus problem in synchronous and asynchronous executions] \label{prop:synclb} Assume that messages carry a sender and/or a receiver ID. Assume also that the consensus algorithm terminates within $R$ synchronous rounds and that that agents' UIDs are selected from a set $S$ of cardinality $|S|\geq R (n+1)$. Then, for a given consensus function $f$ and class of graphs $\mathcal G$ with $n$ nodes, $\textrm{FC}(f, \mathcal G) \in \Omega((n\log n+b) /d)$. \end{proposition} \begin{proof} In the synchronous setting, if an agent has received no messages from its neighbors, it decides whether to send a message or wait until it the next round based on (i) its UID $i$, (ii) its initial condition $x_i$ and (iii) the number of rounds $r$ elapsed since execution started. For each possible UID $i$, we call $\rho_i$ the \emph{smallest} round such that, for \emph{some} initial condition $x_{(i,\rho_i)}$, a node with UID $i$ and initial condition $x_{(i,\rho_i)}$ sends a message at round $\rho_i$ if it has not received a message until round $\rho_i-1$. Informally, $\rho_i$ is the first round when a node with UID $i$ may (for at least one initial condition) send a message if it hasn't received one; $\rho_i-1$ is the last round when a node with UID $i$ can not send a message unless it has received one, irrespective of its initial condition. The number $\rho_i$ can only assume $R+1$ possible values (including $\infty$ if the node never sends a message before round $R+1$ unless it has received one). Furthermore, at most $n-1$ UIDs can have $\rho_i=\infty$: otherwise an adversary could build a network where no agent sends a message before round $R+1$, and therefore consensus is not achieved is achieved within $R$ rounds. Then, by the pigeonhole principle, there exists a round $\bar \rho$ with $1\leq \bar\rho \leq R$ such that at least $n$ UIDs have $\rho_i=\bar \rho$. An adversary can arrange agents with these UIDs (and matching initial conditions) in a network: then all $n$ agents will send a message at round $\bar \rho$, after staying silent for the first $\bar \rho -1$ rounds. Since every message carries the transmitter or the receiver UID, the size of each message is lower-bounded by $\log n$. Thus, $n\log n$ bits may be sent at round $\bar\rho$. Our argument is completed by the observation that, as in Proposition \ref{prop:asynclb}, every agent sends its initial value at least once: this operation has a bandwidth complexity of $\Omega((\log n + b)/d)$. Thus, the broadcast complexity of the consensus problem in the synchronous case is $\textrm{FC}(f, \mathcal G) \in \Omega((n\log n +b)/d)$. \end{proof} We have now proven a lower bound on the bandwidth complexity of the consensus problem for synchronous and asynchronous executions. In the following section we show tightness of this bound by proposing a bandwidth-optimal algorithm, then we compare its performance with commonly-used consensus procedures. \section{Tightness of the lower bound and bandwidth complexity of common consensus algorithms} \label{sec:algs} In this section, we analyze the byte complexity of the average-based consensus algorithm \cite{JNT:84-extra,ROS-RMM:03c,AJ-JL-ASM:02}, the flooding consensus algorithm \cite{NL:96} and the GHS algorithm with convergecast \cite{RGG-PAH-PMS:83,FR-MP:13}. We also propose a modified version of the GHS algorithm that achieves the lower bound presented in Propositions \ref{prop:asynclb} and \ref{prop:synclb}. \subsection{Bandwidth complexity of common consensus algorithms} \begin{lemma}[Bandwidth complexity of the average-based consensus algorithm] \label{prop:avgbased} Assume that messages carry a sender and/or a receiver UID. Then the bandwidth complexity of the average-based consensus algorithm is $O(n(\log n + b)/d)$, where $b$ is the size of an agent's initial condition. \end{lemma} \begin{proof} In the average-based consensus algorithm, nodes maintain a local estimate of the consensus value (which has the same type and size as the nodes' initial value). At each time step, nodes send their estimate to their neighbor, then update their estimate as the \emph{average} of their estimate and their neighbors'. Thus, if messages carry the sender UID, each node transmits $\log n +b$ bits with each message. All agents may communicate at once: this is the case, for instance, in synchronous executions, which are a special case of more general asynchronous executions. The resulting bandwidth complexity is $O(n(\log n + b)/d)$. \end{proof} \begin{lemma}[Bandwidth complexity of the flooding consensus algorithm] \label{prop:flooding} Assume that messages carry a sender and/or a receiver UID. Then the bandwidth complexity of the flooding consensus algorithm is $O(n^2(\log n + b)/d)$, where $b$ is the size of an agent's initial condition. \end{lemma} \begin{proof} In a flooding algorithm, at each time step, every node sends to all neighbors all \emph{new} information it has received at the previous (asynchronous) round. It is easy to observe that, for certain network topologies (e.g. for the complete graph) every node may receive information from $O(n)$ other nodes at the same time and thus retransmit information from $O(n)$ nodes simultaneously. Each piece of information from one node has size $\log n + b$: thus, all nodes may send messages of size $n(\log n + b)$. Furthermore, all nodes may send large messages simultaneously (e.g., in a synchronous execution). Thus the bandwidth complexity of the flooding algorithm is $O(n^2(\log n + b)/d)$. \end{proof} \begin{lemma}[Bandwidth complexity of the GHS algorithm with convergecast] \label{prop:GHSub} Assume that messages carry a sender and/or a receiver UID. Assume also that the consensus function is hierarchically computable. Then the bandwidth complexity of the GHS algorithm with convergecast is $O((n\log n+nb)/d)$, where $b$ is the size of an agent's initial condition. \end{lemma} \begin{proof} The GHS algorithm builds a rooted minimum spanning tree by repeatedly merging non-spanning trees across (minimum-weight) edges. The algorithm requires each edge to have a unique weight, while our model considers an unweighed graph: thus, we assign to each edge a unique, arbitrary, weight\footnote{One popular choice for edge weights is to assign to each edge a ``weight'' equal to the UIDs of the two nodes incident on the edge and use a lexicographic ordering.}. Messages exchanged by the agents during execution only contain node IDs, edge weights and boolean values: thus, the size of each message is upper-bounded by $O(\log n)$. Since no message is larger than $\log n$ and nodes only send one (broadcast) message at any given time, the bandwidth complexity of the GHS algorithm is $O(n\log n /d)$. We refer the interested reader to \cite{RGG-PAH-PMS:83} for an in-depth discussion of the algorithm. Once a rooted tree has been established, the root contacts all other nodes via a tree broadcast and asks them to relay their values. Nodes then relay their values through a tree convergecast: every node waits until it has heard back from all its children (if any), then computes the consensus value for its branch (this is always possible if the consensus function is hierarchically computable) and relays it to its parent. When the root learns every child's consensus value, it computes the overall consensus value and relays it to every node via a tree broadcast. All messages exchanged in this phase have size $O(\log n + b)$: every message contains a consensus value of size $b$ and a sender and/or a receiver ID. There exist tree network topologies (e.g. a tree of depth one) where up to $n-1$ nodes may send messages simultaneously during a tree convergecast: thus, the bandwidth complexity of this phase is $O(n(b+\log n)/d)$. This completes our proof. \end{proof} We remark in passing that Awerbuch's minimum spanning tree algorithm \cite{BAw:87}, which improves the GHS algorithm's time complexity to a time-optimal $O(n)$, is \emph{not} bandwidth-optimal: in particular, the \emph{Test-Distance} procedure has a bandwidth complexity of $O(n\log^2 n)$ during certain executions on selected network topologies, since each Test-Distance message needs to keep track of the return route to its sender. \subsection{Tightness of the lower bound on bandwidth complexity} The GHS-inspired consensus algorithm outlined above is not bandwidth-optimal: optimality, however, can be achieved with a simple modification that does not influence asymptotic time, message or byte complexity\footnote{Informally, time complexity is the time required by the algorithm to converge, message complexity is the overall number of messages exchanged by all agents and byte complexity is the overall number of bytes exchanged by all agents. We refer the reader to \cite{FR-MP:13} and \cite{FR-MP:14a} for a rigorous definition of these complexity metrics.}. The tree-building phase is unchanged. When computing the nodes's consensus function, we exploit the tree structure to make sure that only one node sends a message at any given time. Specifically, once a rooted tree has been established, every node contacts its children \emph{one by one}: children are ordered arbitrarily and every child node is only contacted once the previous child has returned its branch's consensus function. Pseudocode for the algorithm is reported in Algorithm \ref{alg:slowconvergecast}. \begin{algorithm} \caption{Bandwidth-optimal consensus function computation on a tree} \label{alg:slowconvergecast} \begin{algorithmic} \floatname{algorithm}{Procedure} \renewcommand{\algorithmicrequire}{\textbf{Input:}} \renewcommand{\algorithmicensure}{\textbf{Output:}} \Require $IsRoot$ \Comment{a bool} \State $Children$ \Comment{An ordered list of child nodes} \State $Parent$ \Comment{The parent node's ID} \State $x$ \Comment{The node's initial condition} \State $c$ \Comment{A local estimate of the consensus value} \If{$IsRoot$ is true} \State\Call{ComputeConsensusValue}{}\Comment{The root initializes the consensus procedure} \EndIf \Procedure{ComputeConsensusValue}{} \ForAll{$Child$ in $Children$} \State Ask $Child$ to \Call{ComputeConsensusValue}{} \State Wait for $Child's$ \Call{ConsensusReply}{$c_{Child}$}. \State $c \gets$ \Call{UpdateConsensus}{$c$,$c_{Child}$} \EndFor \If{$IsRoot$ is true} \Comment{Inform every node of the consensus value} \State Ask $Child$ to \Call{RelayConsensusValue}{c} \State Wait for $Child's$ \Call{AckConsensus}{}. \Else \State Send $Parent$ a \Call{ConsensusReply}{$c$}\Comment{Inform parent of the branch's consensus value} \EndIf \EndProcedure \Procedure{RelayConsensusValue}{$c_r$} \State $c\gets c_r$\Comment{Store the consensus value} \ForAll{$Child$ in $Children$} \State Ask $Child$ to \Call{RelayConsensusValue}{c} \State Wait for $Child's$ \Call{AckConsensus}{}. \EndFor \If{$IsRoot$ is false} \State Send $Parent$ an \Call{AckConsensus}{} \EndIf \State Terminate \EndProcedure \Function{UpdateConsensus}{$c_i,x_j$} \State $x_i\gets c_i * x_j$ \EndFunction \end{algorithmic} \end{algorithm} It is easy to prove that the bandwidth complexity of this consensus function computation on a tree is $O((\log n + b)/d)$. \begin{lemma}[Bandwidth complexity of Algorithm \ref{alg:slowconvergecast}] \label{prop:ccfc} Assume that messages carry the sender and/or the receiver ID. Assume also that the consensus function is hierarchically computable. Then the bandwidth complexity of Algorithm \ref{alg:slowconvergecast} is $O((\log n + b)/d)$. \end{lemma} \begin{proof} The algorithm is, essentially, a token-passing algorithm. The token originates at the root. Nodes, starting with the root, pass the token to their children, one at a time, when they ask each child to compute its consensus value; children pass the token back to their parent when they relay the local estimate of the consensus value. Nodes (starting at the root) then pass the token to their children, one at a time, when they relay the consensus value; children return the token to their parent when they acknowledge reception of the consensus value. It is easy to see that (i) a node only sends a message when it holds the token and (ii) the token is never duplicated, since each node only contacts one child when it has heard back from the previous child and each node only has one parent. Thus, only one node can send a message at any given time. Messages contain a local estimate of the consensus value, a sender ID and a receiver ID: their size is $b+2\log n$. The claim follows. \end{proof} Lemma \ref{prop:ccfc} allows us to prove tightness of the lower bounds on bandwidth complexity. \begin{proposition}[Bandwidth complexity of the consensus problem in asynchronous executions] Assume that messages carry the sender and/or the receiver ID. Assume also that agents' UIDs are selected from a set $S$ of cardinality $|S|\geq 2n$ and that the consensus function is hierarchically computable. Then, for a given consensus function $f$ and class of graphs $\mathcal G$ with $n$ nodes, $\textrm{FC}(f, \mathcal G) \in \Theta((n\log n+b) /d)$. \end{proposition} \begin{proof} In order to prove the claim, we need to find an algorithm with bandwidth complexity $O((n\log n +b)/d)$. The GHS algorithm has a bandwidth complexity of $O(n\log n)$, as shown in the first part of Lemma \ref{prop:GHSub}. Once a tree structure has been established, Algorithm \ref{alg:slowconvergecast} computes the consensus function with a bandwidth complexity of $O((\log n + b)/d)$. The claim then follows from Proposition \ref{prop:asynclb}. \end{proof} The proof of the next proposition is identical to the one above and is therefore omitted. \begin{proposition}[Bandwidth complexity of the consensus problem in synchronous and asynchronous executions] Assume that messages carry a sender and/or a receiver ID. Assume also that the consensus algorithm terminates within $R$ synchronous rounds and that that agents' UIDs are selected from a set $S$ of cardinality $|S|\geq R (n+1)$. Then, for a given consensus function $f$ and class of graphs $\mathcal G$ with $n$ nodes, $\textrm{FC}(f, \mathcal G) \in \Theta((n\log n+b) /d)$. \end{proposition} We note in passing that Algorithm \ref{alg:slowconvergecast} has the same worst-case time, message and byte complexity as the convergecast algorithm proposed in \cite{FR-MP:13}, as shown in the two lemmas below. \begin{lemma}[Time complexity of Algorithm \ref{alg:slowconvergecast}] \label{prop:cctc} The time complexity of Algorithm \ref{alg:slowconvergecast} is $O(n)$. \end{lemma} \begin{proof} Every node except for the root receives two messages from its parent (one asking to compute its branch's consensus value and one relaying the network's consensus value) and sends two messages to its parent (one informing the parent of the branch's consensus value and one to acknowledge reception of the network's consensus value). Only one message is sent at any given time,as shown in Proposition \ref{prop:ccfc} . Thus, the time complexity of Algorithm \ref{alg:slowconvergecast} is $4(n-1)$. \end{proof} \begin{lemma}[Message and byte complexity of Algorithm \ref{alg:slowconvergecast}] Assume that messages carry the sender and/or the receiver ID. Assume also that the consensus function is hierarchically computable. Then the message and byte complexity of Algorithm \ref{alg:slowconvergecast} are $O(n)$ and $O(n(\log n + b)$ respectively. \end{lemma} \begin{proof} As shown in Proposition \ref{prop:cctc}, $4(n-1)$ messages are sent in the network. Furthermore, if the consensus function is hierarchically computable, each message has size $(\log n + b)$, as shown in Proposition \ref{prop:ccfc}. The claim follows. \end{proof} \section{A tunable algorithm} \label{sec:tunable} The algorithms presented in Section \ref{sec:algs} offer optimal performance with respect to different complexity metrics.\ The flooding algorithm is time-optimal for any graph (and not only worst-case optimal over the class $\mathcal G$ of graphs with $n$ nodes); furthermore, it offers \emph{maximal} robustness to disruptions of a communication channel and performs well on time-varying networks. On the other hand, the bandwidth complexity of flooding is the worst among the algorithms we study and its byte complexity (discussed in \cite{FR-MP:13}) is also very suboptimal. The average-based algorithm performs no better than flooding with respect to byte complexity and robustness, and it has significantly worse time complexity. On the other hand, its bandwidth complexity is better than the flooding algorithm's, but it is not optimal. The GHS-inspired algorithm has optimal bandwidth and byte complexity. However, the algorithm has minimal robustness margins to the disruption of a communication channel. In this section we show how the hybrid, tunable algorithm proposed by the authors in \cite{FR-MP:13} achieves \emph{intermediate} bandwidth performance between the flooding algorithm and the GHS-inspired algorithm, recovering the time-optimal behavior of the flooding algorithm and the bandwidth-optimal behavior of GHS for different values of the tuning parameter. This result complements our previous findings on the time and byte complexity of the hybrid algorithm and shows how this algorithm can achieve \emph{mixed} time/bandwidth performance metrics. Due to space limitations, we only report a high-level description of the algorithm: we refer the interested reader to \cite{FR-MP:13} for details\footnote{We remark two very minor changes with respect to the detailed description in \cite{FR-MP:13}: we replace the challenge-response exchanges at the end of Phase 1 and at the beginning of Phase 3 with broadcasts, to better exploit the \emph{local broadcast} communication model; furthermore, we do not duplicate messages exchanged between clusters for simplicity.}. Our algorithm operates in four phases. Phase 1 starts by building a forest of \emph{minimum weight} trees (shown in Figure \ref{fig:hybridalgphase1}) of height $O(n/m)$. All nodes run a modified version of the GHS algorithm \cite{RGG-PAH-PMS:83} which only differs from the original algorithm with respect to the stopping criterion. When a node discovers that Phase 1 is over, it informs its neighbors with a broadcast. When a node has finished Phase 1 and all its neighbors have, too, it enters Phase 2. In Phase 2, tree height is upper-bounded by splitting clusters with too many agents while enforcing a lower bound on tree size. This phase of the algorithm starts at the leaves of each tree. Each node recursively counts the number of its descendants moving towards the root; agents with more than $\lfloor n/m \rfloor$ offspring create a new cluster, of which they become the root, and cut the connection with their fathers. The tree containing the original root may be left with too few nodes: the root can undo one cut to counter this. In Phase 3, each tree establishes connections with neighbor clusters, as shown in Figure \ref{fig:hybridalgphase3}. When a node switches to Phase 3, it informs all neighbors of its Cluster ID with a broadcast. The root of each cluster is then informed of the available connections to neighbor clusters with a convergecast starting at the leaves. Specifically, as soon as a node knows (i) what clusters its children are connected to (either directly or through their children) and (ii) each neighbor agent's cluster ID, it informs its parent about which clusters it is connected to (either directly or indirectly). Roots also exploit the tree structure to compute their cluster's consensus function via Algorithm \ref{alg:slowconvergecast}: once they have computed the cluster's consensus function, they switch to Phase 4. In Phase 4, cluster roots communicate with each other through the connections discovered in the previous stage, as shown in Figure \ref{fig:hybridalgphase4}. Conceptually, this phase of the algorithm is simply flooding across clusters. Each root sends a message containing its cluster's consensus function to each neighbor tree through the connections built in Phase 3. When a root learns new information, it forwards it once to its neighbor clusters via the same mechanism. If a link failure breaks one of the trees (as in Figure \ref{fig:hybridalgphaseF}), the two halves evaluate their size. If either of the two halves is too small, it rejoins an existing cluster; a splitting procedure guarantees that tree height stays bounded. After failure, all nodes update their routing tables. Finally, when a link outside a tree fails, nodes on the two sides of the failure update their routing tables and notify their cluster roots. \begin{figure}[h] \centering \begin{subfigure}[b]{0.2\textwidth} \includegraphics[width=\textwidth]{Phase1.pdf} \caption{Phase 1 and 2: forest building} \label{fig:hybridalgphase1} \end{subfigure} \begin{subfigure}[b]{0.2\textwidth} \includegraphics[width=\textwidth]{Phase3.pdf} \caption{Phase 3: establishment of inter-cluster links} \label{fig:hybridalgphase3} \end{subfigure} \begin{subfigure}[b]{0.2\textwidth} \includegraphics[width=\textwidth]{Phase4.pdf} \caption{Phase 4: inter-cluster flooding} \label{fig:hybridalgphase4} \end{subfigure} \begin{subfigure}[b]{0.2\textwidth} \includegraphics[width=\textwidth]{PhaseF.pdf} \caption{Phase F: recovery from failure} \label{fig:hybridalgphaseF} \end{subfigure} \caption{Schematic representation of the hybrid algorithm behavior.} \label{fig:hybridalgschematic} \end{figure} We now study the bandwidth complexity of the tunable consensus algorithm. \begin{lemma}[Bandwidth complexity of Phase 1 of the tunable algorithm] \label{lemma:hybrid1ub} Assume that messages carry the sender UID. Assume also that the consensus function is hierarchically computable. Then the bandwidth complexity of Phase 1 of the tunable algorithm is $O(n\log n/d)$. \end{lemma} \begin{proof} Phase 1 only differs from the GHS algorithm, whose bandwidth complexity is shown to be $O(n\log n/d)$ in Lemma \ref{prop:GHSub}, with respect to the stopping criterion. When a node stops, it informs its neighbors: the size of the message informing the neighbors is $O(\log n)$, since it only contains the node's UID and a constant-size message. All nodes may stop and inform their neighbors at once: thus the bandwidth complexity of Phase 1 is $O(n\log n /d)$. \end{proof} \begin{lemma}[Bandwidth complexity of Phase 2 of the tunable algorithm] \label{lemma:hybrid2ub} Assume that messages carry the sender UID. Assume also that the consensus function is hierarchically computable. Then the bandwidth complexity of Phase 2 of the tunable algorithm is $O(n\log n/d)$. \end{lemma} \begin{proof} The size of all messages exchanged in Phase 2 is $O(\log n)$: each message contains the number of agents that are descendants of the sender node and a cut ID containing two UIDs. Every node has at most $n-1$ descendants: thus, the number of descendants of a given node can be represented with $O(\log n)$ bits. Since every node sends at most one message at a time, the bandwidth complexity of Phase 2 is $O(n\log n/d)$. \end{proof} \begin{lemma}[Bandwidth complexity of Phase 3 of the tunable algorithm] \label{lemma:hybrid3ub} Assume that messages carry the sender UID. Assume also that the consensus function is hierarchically computable. Then the bandwidth complexity of Phase 3 of the algorithm is $O((nm\log n+b)/d)$. \end{lemma} \begin{proof} Two types of messages are exchanged (potentially at the same time) in the cluster discovery routine in Phase 3. First, nodes announce their cluster ID: each message has size $\log n$ (since a cluster ID is the ID of the root node) and every node sends exactly one such message. Then, nodes send their parents a list of the clusters their branch is connected to. Every message contains up to $m$ cluster IDs, each of size $\log n$: thus, the size of each message is upper-bounded by $m\log n$. Every node sends at most one message at any given time: thus the bandwidth complexity of Phase 3 is $O(nm\log n/d)$. In addition, each root computes the cluster's consensus function with Algorithm \ref{alg:slowconvergecast}: the bandwidth complexity of this operation is $O((\log n +b)/d)$, as shown in Lemma \ref{prop:ccfc}. Thus, the overall byte complexity of Phase 3 of the algorithm is $O((nm\log n+b)/d)$. \end{proof} \begin{lemma}[Bandwidth complexity of Phase 4 of the tunable algorithm] \label{lemma:hybrid4ub} Assume that messages carry the sender UID. Assume also that the consensus function is hierarchically computable. Then the bandwidth complexity of Phase 3 of the algorithm is $O(nm(b+\log n)/d)$ and $O(m^3(b+\log n)/d)$ . \end{lemma} \begin{proof} We prove the two bounds separately. First, we note that that every message exchanged in Phase 4 contains a list of at most $m-1$ intermediate consensus values (one per cluster) and the ID of the relevant cluster: thus, the size of each message is $O(m(\log n + b))$. Since no node sends more than a message at a time, the first bound follows. Second, we observe that in the cluster flooding algorithm every cluster forwards each new piece of information it receives from a (neighbor or non-neighbor) cluster to neighbor clusters exactly once. Thus, each cluster contacts each of the $O(m-1)$ neighbor clusters with at most $m(b+\log n)$ bits of information overall. Each exchange between neighbor clusters may require multiple messages. However, it is easy to see that only one message per exchange is transmitted at a given time: the message is \emph{routed} from the root of a cluster to the root of its neighbor, thanks to the routing tables developed in Phase 3, with no duplication. Thus, even if all $m$ clusters send information about all $m-1$ other clusters to every neighbor at the same time, no more than $O(m^3(\log n + b)/d)$ bits are exchanged at any given time. Once a cluster root has computed the overall consensus value, it informs all nodes in its cluster. This is done with Algorithm \ref{alg:slowconvergecast}, with bandwidth complexity $O((\log n + b)/d)$. The overall bandwidth complexity of Phase 4 is therefore also upper-bounded by $O(m^3(\log n + b)/d)$. \end{proof} \begin{proposition}[Bandwidth complexity of the tunable algorithm] The bandwidth complexity of the tunable algorithm is $O(m^3(b+\log n)/d)$ and $O(nm(b+\log n)/d)$. \end{proposition} \begin{proof} The proof follows immediately from Lemmas \ref{lemma:hybrid1ub}, \ref{lemma:hybrid2ub}, \ref{lemma:hybrid3ub} and \ref{lemma:hybrid4ub}. \end{proof} \section{Conclusions} \label{sec:conclusions} In this paper we study the bandwidth complexity of the consensus problem on networks with omnidirectional communication, with particular attention to \emph{robotic} applications. We provide a novel definition of bandwidth complexity which captures the bandwidth use of multi-agent systems with modern Media Access Control mechanisms. This definition allows us to show that, even in network of moderate size, bandwidth use can be a limiting factor on the time performance and on the scalability of common consensus algorithms such as flooding. We then prove a lower bound on the bandwidth complexity of the consensus problem that becomes tight for hierarchically computable consensus functions and we provide a matching bandwidth-optimal algorithm. Finally, we extend our previous results in \cite{FR-MP:13}, proving that the hybrid algorithm presented in the paper achieves intermediate bandwidth complexity between the lower bound and the bandwidth complexity of the time-optimal algorithm, according to a user-defined tuning parameter. The tradeoff between worst-case time performance, byte performance and bandwidth performance is shown in Figure \ref{fig:bounds}. The implication of this result is that the hybrid algorithm can be used to achieve \emph{mixed} performance metrics, trading time performance and robustness for byte complexity (representative of energy consumption for communication) and bandwidth complexity. We conclude this paper with a discussion of the limitations of our analysis, which reflect in interesting directions for future research. First, our worst-case analysis provides lower bounds on bandwidth complexity for the class $\mathcal{G}$ of \emph{all} graphs with $n$ nodes: it is of interest to further refine our results to capture the effect of network topology (and in particular of the maximum node degree) on the fundamental limitations on bandwidth performance and on the performance of existing algorithms. Second, an average-case analysis over graphs and asynchronous executions drawn randomly from a representative probability distribution would provide significant insight into the effect of bandwidth complexity on real-world networks, where a (small) probability of collisions is acceptable in presence of a TCP mechanism. Third, while the class of hierarchically computable functions encompasses many relevant engineering problems, it is of interest to study the role of bandwidth complexity for consensus functions that are \emph{not} hierarchically computable. In particular, nonhierarchical algorithms such as average-based consensus may perform no worse than hierarchical algorithms such as GHS with convergecast when the consensus function does not benefit from hierarchical computation. Finally, accurate software and hardware simulations of the performance of the algorithms presented in this paper on wireless channels with different MAC mechanisms will provide further insight into the relevance of our bandwidth metric and into the relative benefits of the different approaches. \bibliographystyle{IEEEtran}
\section{Introduction} Few would disagree that fostering scientific literacy among the general public is a worthwhile goal. We live in a world of increasing technological complexity, and developments in biotechnology, energy use, communications, and many other fields impact people's lives in unprecedented ways. Moreover, the advance of science has illuminated countless fascinating aspects of the inner workings of nature, from the structure of stars to the interactions of genes, and an understanding of science opens the doors to an enriching understanding of these insights. There is widespread concern, however, that the level of scientific literacy in contemporary society is poor, with respect to both basic scientific knowledge and, more importantly, understanding of the nature of the scientific process \cite{pew2009, miller2004, gross2006}. One way to address this is via general education undergraduate courses --- i.e. courses intended for students not majoring in the sciences or engineering --- which in many cases provide these students' last formal exposure to science. A variety of such courses exist in the Physics departments at many universities, structured as overviews of wide swathes of the subject, or forays into more specialized niches. I describe here a course on \textit{biophysics} for non-science-major undergraduates, titled ``The Physics of Life,'' that I have recently developed and taught at the University of Oregon. Biophysics, I claim, is a particularly useful vehicle for addressing scientific literacy. It involves important and general scientific concepts, demonstrates connections between basic science and tangible phenomena related to health and physiology, and illustrates how scientific insights do not develop along predictable paths, but rather often arise by the creative application of perspectives and tools from disparate fields. Moreover, it highlights the importance of physics in biological research, a view increasingly realized among biologists \cite{ascb2012, visionandchange, hilborn2014}, but not by the general public. Here I describe the design of this one quarter (ten week) course, the specific content of a few of its modules, its use of active learning and evidence-based pedagogy, and aspects of its enrollment and evaluation. The aim of the article, in addition to documenting aspects of this course, is to hopefully help seed similar classes elsewhere, or instances in which biophysical concepts are incorporated into other general education classes. \section{Goals and Topics} The course has three overarching goals: (1) To help students learn how physical principles guide and constrain life. This includes developing a basic understanding of what the biomolecules and biomaterials that make up organisms are, and how their physical properties and interactions govern their function. (2) To improve students' ability to understand quantitative data and models. This goal spans skills such as numerical estimation and grasping the meaning of graphs, including non-standard graphs such as log-log plots (e.g. for biomechanical scaling relationships). (3) To improve students' comprehension of the process by which scientific understanding develops. This encompasses examples of the complex relationships between ``pure'' and ``applied'' science, and of the connections between seemingly disparate fields of scientific study. Particular topics were chosen to contribute to these goals and develop students' scientific literacy. The course is part of the University of Oregon's Science Literacy Program \cite{slp}, which aims to help implement evidence-based pedagogical methods across a range of general education courses spanning several science departments, and to facilitate new classes and new approaches to faculty and student training. \subsection {Macroscopic topics} The ten week term is roughly divided into two halves. The first covers macroscopic topics, with a focus on scaling concepts, i.e. understanding how various physical forces and biomechanical properties scale with organism size, and how this influences the behavior and physiology of animals and plants. \subsubsection{Surface tension} The first biological question we address is why small insects can walk on water, while humans cannot. This leads to the concept of surface tension, introduced by demonstrating a metal paper clip sitting atop a water surface. The simple question, ``Why does it stay up?'' is a surprisingly difficult one to answer; many students will state ``surface tension,'' but when probed will not be able to explain what this means, leading to interesting conversations on the inadequacy of simply naming phenomena as compared to understanding the mechanisms underlying them \cite{feynman}. We then discuss the nature of liquids, and how a consequence of intermolecular attraction is a tendency to minimize surface area, and hence surface tension. There is, of course, a force associated with surface tension, holding up our paper clip against the force of gravity. What geometric properties of the object should this force depend on? With two objects of equal area, but different edge lengths atop a bath of water (Figure~\ref{chopperwheels}), adding weights to each until they sink, one can show quite simply that the shape with the greater perimeter can support considerably more weight, and so has a larger surface tension force associated with it. (It's a surprisingly dramatic demonstration; the students make guesses beforehand and are almost breathless throughout the slow addition of weights.) From this, we establish that the force associated with surface tension scales with length and that, all things being equal, an organism whose dimensions double would, at a liquid surface, experience an upward force that is twice as large. We also learn that the force of gravity is proportional to the mass of an object, and hence scales as the cube of length, which points to the answer to our original question: if we imagine organisms growing in size, the force of gravity increases much more than the force that surface tension can provide. This difference in scaling behavior underlies differences in animal behavior, and explains why we don't find large animals walking on water. It also explains why an individual fire ant, for example, can walk on water due to surface tension but a raft of ants cannot (which students are able to predict, based on their improved physical understanding), which leads ants agglomerating in flooding jungles to trap air bubbles to harness the force of buoyancy to keep themselves up, the subject of recent, fascinating studies \cite{mlot_fireants}. This leads simply to a topic of considerable physiological importance: breathing. The surface of the lungs is wet, and so much of the work necessary for breathing is work done against surface tension. (For this reason it is easier to inflate lungs with water than with air \cite{clements62}, which students are surprised to learn.) Our lungs secrete, therefore, a surfactant that lowers the surface tension of the lungs and facilitates breathing. This surfactant is produced rather late in gestation, however, around week 30, and its absence leads to Infant Respiratory Distress Syndrome (IRDS), the leading cause of death among premature infants. The mortality rate from IRDS has dropped from about 25,000 deaths per year in the 1960s to less than 1000 in 2005 \cite{Schraufnagel}, due to the development of surfactant treatments --- in essence, injecting animal-based or synthetic amphiphilic molecules into the lungs. The connections between a basic physical concept, a biological function, and a real-world application are rarely clearer than this. \begin{figure}[h!] \centering \includegraphics[width=3.5in]{IMG_1006_chopper1.jpg} \caption{Two chopper wheels, with the same area but different contact perimeters, are supported atop water by surface tension. Adding weights to each one until it sinks helps demonstrate that the force associated with surface tension is dependent on perimeter, which leads to an understanding of the scaling properties of interfacial forces.} \label{chopperwheels} \end{figure} Of course, the discussion of surface tension above will seem trivial to most readers of this journal. It serves a useful function in the course, however, beyond being interesting, and that is to introduce scaling concepts. This takes a considerable amount of work --- a statement as simple as ``volume is proportional to length cubed'' is not only foreign to most students but is remarkably non-intuitive. In general, their own prior exposure to geometry has been centered on memorizing formulas for, for example, the volumes of various shapes, rather than developing more expansive notions of concepts like volume and area. I describe later in this article various exercises involving, for example, making log-log plots of area and length for simple shapes, and measuring volumes of complex shapes, that build intuition about geometric scaling relationships. \subsubsection{Biomechanics and scaling} We then examine other issues of biomechanical scaling, especially the question of why larger land animals need disproportionately thicker bones than smaller ones. An elephant's femur, for example, is about 10 times longer than a small dog's, but has a diameter about 20 times greater. It is straightforward to illustrate this with images of bones; alternatively, one can find real bones (Figure~\ref{elephantfemur})\cite{UOelephant}. Why are the bones so disproportionate? Again, the scaling of different physical properties provides an explanation. The force of gravity is proportional to mass, and hence length cubed, while the strength of bones, or beams in general, is proportional to their cross-sectional area and hence the square of length. (One could of course further explore the continuum mechanics of buckling and make the preceding statement more accurate; we do not in this course.) To avoid being crushed, larger animals need disproportionately wider bones. Notably, if the bone diameter scales as length$^{3/2}$, gravitational force and bone strength follow one another; this is the case for ``mechanically similar'' animals. Following an example from McMahon and Bonner \cite{mcmahon1983}, we examine plots of bone dimensions for a wide range of bovids (antelope, wildebeest, etc.), and find that this mechanical similarity holds, highlighting a non-obvious shared characteristic of these diverse animals. The topic of bone shape and scaling has a long history, dating at least to Galileo \cite{galileo}, and is compellingly discussed in a variety of books (e.g. Refs.~\onlinecite{mcmahon1983, lifesdevices}), discussed further below. \begin{figure}[h!] \centering \includegraphics[width=3.5in]{students_and_elephant_bone.jpg} \caption{An elephant femur helps illustrate the biophysics of bone shape, and attracts attention while being carted through campus.} \label{elephantfemur} \end{figure} In addition to their relevance to animal form, the biomechanical scaling ideas explored above relate to contemporary issues of human health in intriguing ways. The Body Mass Index (BMI), for example, postulates a person's mass ($M$) divided by height squared ($h^2$) as a convenient and size-invariant measure of obesity. If people of different heights were the same shape (which is not the case), one would expect $m/h^3$ to be a useful measure. The BMI assumes a particular non-isometric form, with $m \sim h^p$ and $p=2$. Do actual data on masses and heights obey this? Strikingly, they do not \cite{korevaar}; to the extent that there is power-law scaling at all, $p$ is roughly 2.6-2.7. The peculiarities of the BMI, such as the inaccuracy of its obesity implications for tall people, are familiar to many students, which leads to interesting discussions of why the measure exists and persists. \subsection {Microscopic topics} The second half of the course covers microscopic topics, especially the cellular and subcellular phenomena that are the targets of most contemporary biophysical study. A key goal is to convey an understanding of random, Brownian motion and its importance. Students are used to seeing cartoons or diagrams of biological processes, and from these form the mistaken impressions that these processes are more ordered and deterministic than they actually are, and that small-scale ``machines'' can be thought of simply as scaled-down versions of macroscopic devices. Reality is, of course, much different, and seeing how dissimilar from our familiar experience the microscopic world is enhances our appreciation of it. \subsubsection{Brownian Motion} We begin with observations, either with tabletop microscopes or previously recorded videos, of colloidal Brownian motion, noting also its scientific history \cite{mazo}. We introduce the idea of a random walk. To characterize this, rather than constructing an algebraic derivation, we turn again to now-familiar tools for uncovering scaling behavior, plotting various properties of random walks simulated in class and finding, eventually, that the root mean square distance traveled robustly scales as time$^{0.5}$. This non-linear scaling of distance and time, together with a few numbers, explains why small cells, like bacteria, can rely on simple random diffusion to distribute material within them, while larger eukaryotic cells must employ active, directed mechanisms. \subsubsection{Biomolecules} We explore the large-molecule components of cells, DNA, proteins, and lipids, examining especially how their physical attributes are integral to their function. The concept of self-assembly is central, and we examine how the combination of simple physical forces and ubiquitous Brownian motion generates structure. Protein folding provides an important example of this. Building on information about protein sizes and prior exposure to the diffusivity of small molecules, students can estimate the timescale required for a chain of amino acids to explore configuration space and adopt a shape. There are numerous connections between this topic and issues of contemporary interest, even beyond the roles of particular proteins, for example the computational challenges of predicting protein folding outcomes \cite{dill2012}, and the consequences of misfolding in diseases such as mad cow disease and Kuru \cite{pruisner1995}. (The latter, spread by cannibalism, is particularly entertaining to discuss.) Further aspects of protein structure can be explored in group projects, described below. Lipid membranes provide still further opportunity to examine self-assembly, as well as enabling connections to earlier discussions of surface tension and to contemporary research into the mechanical properties of these biological structures \cite{parthasarathycurvature2007, lingwood2010, HonerkampSmith2012}. DNA is the most iconic biomolecule, and we examine some of the physics related to its role as a conveyor of genetic information. The packaging issues associated with DNA are easy to introduce. We note that each of us have roughly one meter of DNA in each of our roughly one-micron-diameter cell nuclei. We ask: Is this impressive? Since 1 m is much larger than 1 $\mu$m, an obvious answer is ``yes.'' However, we then ask for a simple estimate of the volume of the nucleus, $\sim (10^{-6}m)^3$, and the volume of the DNA, $\sim 1 \mathrm{m} \times (10^{-9}\mathrm{m})^2$, finding that they are similar, so an equally straightforward answer to our question is ``no.'' Both responses are inadequate, however. To answer meaningfully, we must consider the mechanical properties of DNA. A simple way to do so \cite{pboc} that follows naturally from earlier course topics is to model DNA as a random walk of straight segments, each of length equal to the molecule's persistence length, $\approx 50$ nm. The characteristic size of such a walk, or equivalently the size of a ``blob'' of DNA on its own, is about 200 $\mu$m, showing that its packaging inside the nucleus is, indeed, impressive. Physics highlights the remarkable challenges involved in packing DNA, and also illuminates the tactics employed in response: DNA is highly negatively charged, and positively charged histone complexes serve as spools on which DNA is wound. The topic of DNA again connects with issues of scaling, and also links to contemporary studies on, for example, the even denser packing of DNA in many viruses \cite{evilevitch2003}, and the feedback between DNA packaging and the genetic code \cite{segal2006}. Moreover, it highlights the process of model construction in science, and leads to discussions of the motivations, the limitations, and the utility of models. \subsection {Other topics} The themes of the course offer abundant opportunities for extensions and personalization of topics, incorporating modern insights into entropy in biomolecular systems, pattern formation, energy flows, experimental tools, and countless other topics. For example, in some terms we have explored the fundamentally different ways in which large and small organisms must propel themselves in fluids, \cite{vogelfluids, purcell1977}, a topic that transcends the microscopic and macroscopic divide. This again serves to illustrate that many living creatures inhabit a strange and alien world, at odds with the intuition we develop as large animals in turbulent surroundings. \section{Components of the Course} The non-standard subject matter of the course and its audience of non-science-major undergraduates, who are in general rather averse to mathematics, present challenges for teaching that I have attempted to address through the development of a variety of course materials and activities. Most class sessions involve a small amount of time spent lecturing, with the considerable majority of the period devoted to active learning in one of several forms. In general, the benefits of active learning methods on student performance are increasingly well appreciated for introductory courses for science majors \cite{deslauriers2012, smith2009, freeman2014}. For general education courses this is has been much less explored, but I would argue that active learning is even more important in this context. Engagement with the material is critical, and since the students in general are less interested in science than are science majors, reliance on passive absorption of new concepts and techniques is not very effective. In addition, students are often trained by prior experiences to believe that they are incapable of scientific inquiry. Making a large fraction of the course require the construction of questions, discussion with peers, and other activities not only helps address this, but does so in a way that makes it clear that this way of learning is the expected norm for the course. The most significant tools to aid active learning that we have employed are ``clicker''-based questions about scientific concepts, graphs, or in-class demonstrations\cite{beatty2006}, and in-class worksheets. I have found the worksheets to be highly effective. In these, particular lessons are broken down into a series of questions and discussion topics that students work on in small groups, while teaching assistants and I walk through the class offering advice and asking questions. After most groups have answered a few of the questions, or if many groups are stuck, we all reconvene to go over the topic. For example, our worksheet on the physics underlying bone dimensions began with an exercise plotting bone length and diameter on a logarithmic graph, asked students to sketch graphs that would correspond to isometric- and mechanically-similar scaling, continued with further questions, and finally concluded with a question that encapsulates these concepts and leads to discussion: ``Why can't elephants jump?'' (A few sample worksheets are provided as supplementary materials to this paper \cite{worksheets}.) There is no textbook that spans the range of subjects described above. Assigned readings of excerpts from Steven Vogel's excellent books on biomechanics \cite{lifesdevices, vogelcats, vogelfluids, vogelglimpses, vogelcircuits}, especially \textit{Life's Devices}\cite{lifesdevices}, and McMahon and Bonner's \textit{On Size and Life} \cite{mcmahon1983}, are useful for the macroscopic half of the term. (The two named books also inspired the creation of the course.) John Tyler Bonner's \textit{Why Size Matters: From Bacteria to Blue Whales}\cite{bonnersizematters} is also elegant and clear, and D'Arcy Thompson's classic \textit{On Growth and Form}\cite{OnGrowthandForm} contains innumerable inspiring morphological discussions. I have supplemented excerpts from books with short articles from \textit{Scientific American}\cite{basu2007, clements1962}, \textit{Physics Today}\cite{west2004}, and other sources, as well as materials I wrote myself. The microscopic half of the course relies much more on readings I have written, and brief excerpts from the publicly available \textit{Molecular Biology of the Cell} textbook \cite{mboc}. A list of assigned readings is supplied as Supplementary Material \cite{readinglist}. More than half of the class sessions had an assigned prior reading, with a short quiz on its contents at the start of the period. In addition to readings directly related to class topics, students were given three assignments, to be completed in small groups, in which they read and responded to ``popular science'' articles from \textit{The New York Times}, \textit{The Economist}, and other sources. These dealt with subjects that intersected with those covered in class, for example on scaling relationships claimed to be obeyed by cities, the creation of synthetic nucleotides for DNA, etc. For each article chosen, students were directed to briefly summarize the article, especially the scientific motivations of the work described; to ask one ``quantitative'' question that was not presented in the article or suggest something that could be graphed that would be insightful; and to comment on relationships between the article and in-class topics. Especially since the students in the class are non-science majors, their interaction with science in the future is likely to be largely via popular media of various sorts, and so developing practice with thoughtfully examining popular articles is valuable. We also use more standard assignments and assessments: weekly homework assignments and exams. These focus especially on conceptual understanding of the material and order-of-magnitude numerical estimates. The course has also incorporated a final project in which students, in groups, research some protein with the goal of explaining the relationship between its structure and function. In addition, students also 3D printed a physical model of their protein and compared the usefulness of this visualization with computational rendering (using the widely used PyMol software). (Interestingly, nearly everyone preferred, overall, the computational illustration.) One could easily imagine a greater focus in the course on scientific visualization methods. \section{Challenges and Outcomes} ``The Physics of Life'' has been offered at the University of Oregon four times since 2011, each time with an enrollment of about 60. As intended, students represented many different majors (45), and the majority (78\%) were not science majors. There has been a roughly 2:1 ratio of social science to humanities majors. Of students in the natural sciences, the largest contingent (48\%) were psychology majors, and 10\% (3\% of the total students) were physics majors. The course, therefore, succeeded in its aim of reaching a large number of students not pursuing degrees in the sciences. Student reaction to the course and subject matter has been enthusiastic. In written end-of-term comments, many students have noted the ``incredibly interesting and diverse'' topics, and have ``enjoyed how the material we were learning about related to our everyday lives.'' The University of Oregon Science Literacy Program has been surveying attitudes toward and perceptions of science among students in this and other courses; the results will be documented in the near future \cite{slppaper}. Course evaluations, which tabulate responses to a standard set of university-wide questions, do not provide a meaningful measure of student learning, but are unfortunately the only tool available for inter-course comparisons. Evaluation scores for ``The Physics of Life'' are slightly higher than average for general-education courses offered by the University of Oregon Physics Department (``The Physics of Light and Color,'' ``The Physics of Sound and Music,'' and several others). In the evaluation category of overall course quality, for example, the most recent evaluation score was 4.1 out of 5.0, with the departmental mean and standard deviation for the last 76 general-education courses taught being 3.9 $\pm$ 0.3. Though the course is in general well liked, it presents challenges for students. Many find the mathematical concepts introduced in it difficult. Though nominally simpler than the basic skills in algebra they all have, techniques such as numerical estimation, adeptness with exponents, etc., move beyond their prior habituation with rote memorization of formulae, and require developing a deeper understanding of quantitative perspectives that is non-trivial. For example: via discussion as well as a ``diagnostic'' math quiz during the first week of the term, it is apparent that the considerable majority of students will correctly respond with $x^{a+b}$ when asked what $x^a x^b$ is. However, a much smaller fraction ($\sim 25 \%$), when asked a question like ``if $y$ is proportional to $x^3,$ and $x$ doubles, what happens to $y$?,'' will answer correctly. The former question involves, in most students' experience, memorization of a rule about manipulating symbols; the latter involves an understanding of what exponents mean. Addressing this, while rewarding and ultimately satisfying, takes time. We do a variety of exercises that build from seemingly trivial beginnings, tabulating the volumes and surfaces areas of simple geometric shapes and plotting them versus linear dimensions on logarithmic axes, and measuring the volumes and masses of isometric (same-shape) objects like bolts, establishing how to think about non-linear relationships, and realizing for example that volume is more than the outcome of formulas about shape, but rather that property of space that scales as length cubed. The tangibility of the topics explored in the course, i.e. their applicability to the everyday world of animals and plants, helps students engage with physical concepts. Moreover, connections between the course and physiological relevance are particularly valuable, helping to stimulate interest and appreciation. It is not uncommon for students to have personal experience, for example via family members, with diseases that connect to biophysical properties. (Pre-term births, cystic fibrosis, and cancer have all come up in the course.) Along similar lines, the course serves to illustrate the importance of non-genetic ``information'' in orchestrating life, a message of particular contemporary importance given the tendency of popular media to convey the impression that genes are the sole drivers of function, that there is a gene ``for'' every attribute of health or disease. There is, students learn, no gene that directs lipids into a bilayer, or that ferries neurotransmitters across a chemical synapse; in these and countless other cases, the physical forces and interactions of biomolecules govern and constrain their behavior, a perspective that is important to convey. I will also note that the course is very enjoyable to teach. Being a biophysicist, I am, of course, biased, but having taught various other general education courses in recent years, it is apparent that the variety of the subject matter, its connections to the living world around us, and the contemporary excitement of the field of biophysics all make a general education biophysics course a deeply satisfying and intellectually exciting vehicle with which to convey the message of scientific literacy to a general population. \begin{acknowledgments} The development and implementation of this course have benefitted enormously from its affiliation with the University of Oregon Science Literacy Program (SLP) \cite{slp}, launched with a grant from the Howard Hughes Medical Institute (grant no. 52006956, Science Education Division) and also supported by funds from the University of Oregon. I also gratefully acknowledge input and insights from Elly Vandegrift (associate director of the SLP), Julie Mueller (Univ. of Oregon Teaching Effectiveness Program), Professor Eric Corwin (who taught one term of Physics 171), graduate student assistants Liesl Van Ryswyk, Matt Jemielita, Tristan Hormel, Ryan Baker, Kyle Lynch-Klarup, Savannah Logan, and undergraduate student assistants Kendra Nyberg and Ricky Holton. Undergraduate assistants were upper-division science major supported by the SLP to develop their teaching and communication skills, and to help implement active-learning activities in class. Many the graduate student assistants were wholly or partially supported by SLP. \end{acknowledgments}
\section{Introduction} In this paper we study the operator~$z$ of multiplication by the complex coordinate in Hilbert spaces of holomorphic functions on certain multiply connected domains in the complex plane. The domains we consider are disks with circular holes. The case of a~disk with no holes is the classical one. In the Hardy space of the disk the multiplication operator~$z$ is the unilateral shift whose spectrum is the disk. The $C^*$-algebra generated by the unilateral shift, the Toeplitz algebra, is an extension of the algebra of compact operators by $C(S^1)$, $S^1$ being the boundary of the disk~\cite{Cob}. For the Bergman space the operator~$z$ is a~weighted unilateral shift and its spectrum and the $C^*$-algebra it generates are the same as in the Hardy space~\cite{KL}. Partially for those reasons the Toeplitz algebra is often considered as the quantum disk~\cite{KL,KL2, KL3}. A disk with one hole is biholomorphic to an annulus. In the Bergman space for example, the~$z$ operator is a~weighted bilateral shift with respect to the natural basis of (normalized) powers of the complex coordinate. Its spectrum is the annulus, and the $C^*$-algebra it generates is an extension of the algebra of compact operators~by $C(S^1\times S^1)$, where $S^1\times S^1$ is the boundary of the annulus. The same is true for many other Hilbert spaces of holomorphic functions on an annulus. The resulting $C^*$-algebra is the quantum annulus of~\cite{KL3,KM1}. In this paper we study in detail the two hole case: a~pair of pants. Up to biholomorphism we can realize a~disk with two holes as an annulus centered at zero with outer radius one, with an additional of\/f centered hole. In the space of continuous functions on the closed pair of pants that are holomorphic in its interior, we consider a~specif\/ic inner product with respect to which the operator of multiplication by the complex coordinate~$z$ has a~particularly simple structure. The results we obtain are completely analogous to zero and one-hole cases: the spectrum of~$z$ is the domain of the corresponding pair of pants while the $C^*$-algebra generated by~$z$ is an extension of the algebra of compact operators by $C(S^1\times S^1\times S^1)$, where $S^1\times S^1\times S^1$ is the boundary of the pair of pants. This work is part of an ongoing ef\/fort to understand the structure of quantum Riemann surfaces and their noncommutative dif\/ferential geometry, see~\cite{KL, KL1, KL3, KL4, KM1, KM2,KM3,KM4}. Our paper has many things in common with the work of Abrahamse~\cite{Abr} and Abrahamse--Douglas~\cite{AD}, who use dif\/ferent Hilbert spaces. The paper is organized as follows. Section~\ref{section2} contains an overview of the zero and one-hole cases, while Section~\ref{section3} has a~detailed discussion of the quantum pair of pants. \section{Preliminaries}\label{section2} In this section we describe in some detail, the zero and one-hole cases. Most of the material is well-known, however the treatment of the quantum annulus is somewhat new. \subsection{The quantum disk} In this subsection we look at the structure of the quantum disk. We review the tools and the relevant theorems that will be a~motivation for the subsequent discussion of the quantum pair of pants. Consider the closed unit disk ${\mathbb D}=\{\zeta\in{\mathbb C}: |\zeta|\leq 1\}$. We can represent any holomorphic function inside the disk as a~convergent power series \begin{gather*} f(\zeta)=\sum\limits_{n=0}^\infty e_n\zeta^n. \end{gather*} The Hardy space on the disk is def\/ined as \begin{gather*} H^2({\mathbb D})=\left\{f(\zeta)=\sum\limits_{n=0}^\infty e_n\zeta^n: \sum\limits_{n=0}^\infty|e_n|^2 <\infty\right\}. \end{gather*} We def\/ine the multiplication operator by the complex coordinate, $z:H^2({\mathbb D}) \to H^2({\mathbb D})$ by the formula $f(\zeta)\mapsto \zeta f(\zeta)$. If $E_n=\zeta^n$ is the orthonormal basis on $H^2({\mathbb D})$, then applying~$z$ to the basis elements produces $zE_n=E_{n+1}$ for all $n \geq 0$, i.e.,~$z$ is the unilateral shift; moreover, we have the following formula for the adjoint operator to~$z$ \begin{gather*} z^*E_n= \begin{cases} E_{n-1} &\text{for} \quad n\ge 1, \\ 0 &\text{for} \quad n=0. \end{cases} \end{gather*} Now we consider the $C^*$-algebra generated by~$z$. This well-known algebra is called the Toeplitz algebra, denoted by $\mathcal{T}$, and has also been termed the quantum (noncommutative) disk. This is (partially) based on the following standard results collected here with sketches of proofs which serve as a~guideline for considerations in the next section. \begin{Theorem} The norm of~$z$ is~$1$. The spectrum of~$z$ is all of ${\mathbb D}$, i.e.,~$\sigma(z)={\mathbb D}$. \end{Theorem} \begin{proof} The norm computation is straightforward. By the norm calculation it then follows that the spectrum is a~closed subset of the unit disk. To illustrate that any~$\lambda$ in the interior of ${\mathbb D}$ is an eigenvalue of $z^*$, take $f_\lambda(\zeta) = \sum\limits_{n=0}^\infty \lambda^n \zeta^n$ and so \begin{gather*} z^*f_\lambda(\zeta)=\sum\limits_{n=1}^\infty \lambda^n \zeta^{n-1}=\lambda\sum\limits_{n=1}^\infty \lambda^{n-1}\zeta^{n-1}=\lambda\sum\limits_{n=0}^\infty \lambda^n\zeta^n=\lambda f_\lambda(\zeta).\tag*{\qed} \end{gather*} \renewcommand{\qed}{} \end{proof} Let $\mathcal{K}$ be the algebra of compact operators in $H^2({\mathbb D})$. The next observation tells us how the commutator ideal of $\mathcal{T}$, and $\mathcal{K}$ are related. \begin{Theorem} The commutator ideal of $\mathcal{T}$ is the ideal of compact operators. \end{Theorem} \begin{proof} Since $\mathcal{T}$ is generated by~$z$ and $z^*$, the commutator ideal of $\mathcal{T}$ is equal to the ideal generated by the commutator $[z^*,z]$. Note that $[z^*,z]=P_{E_0}$, the orthogonal projection onto the span of $E_0$. Since this one-dimensional projection is a~compact operator, it follows that the commutator ideal of $\mathcal{T}$ is contained in $\mathcal{K}$. To prove the opposite inclusion we look at the following rank one operators: $E_{ij}(f)=\langle f, E_i\rangle E_j$. Notice that $E_{ij}=z^jP_{E_0}(z^*)^i$, hence those operators belong to the commutator ideal of $\mathcal{T}$. But every compact operator is a~norm limit of f\/inite rank operators, which in turn are f\/inite linear combinations of $E_{ij}$'s. This verif\/ies that the commutator ideal of $\mathcal{T}$ contains $\mathcal{K}$. \end{proof} In order to state the next result, f\/irst we introduce some more notation. We identify $H^2({\mathbb D})$, the Hardy space on the unit disk, with the subspace of $L^2(S^1)$ spanned~by $\{e^{inx}\}_{n\ge0}$. Also given a~continuous function~$f$ on the unit circle, we denote the multiplication operator by~$f$ as~$M_f$. Let $P:L^2(S^1)\to H^2({\mathbb D})$ be the orthogonal projection onto $\spn\{e^{inx}\}_{n\ge0}$, then def\/ine the operator $T_f: H^2({\mathbb D}) \to H^2({\mathbb D})$ by~$T_f=PM_f$. The operator~$T_f$ is known as a~Toeplitz operator. Since $\|M_f\|=\|f\|_{\infty}$, $\|T_f\|\le \|f\|_{\infty}$ and hence it is bounded. We have: \begin{Theorem} \label{toeplitz_disk} The quotient $\mathcal{T}/\mathcal{K}$ is isomorphic to $C(S^1)$, the space of continuous functions on the unit circle. \end{Theorem} \begin{proof} The usual proof constructs an isomorphism between the two algebras. Notice that for a~continuous function~$f$, we have $T_f\in\mathcal{T}$ and since $T_{e^{ix}}$ is the unilateral shift, $T_{e^{-ix}}=T_{e^{ix}}^*$. By the Stone--Weierstrass theorem, every continuous function can be approximated by trigonometric polynomials. Consequently we can def\/ine a~map $\theta: C(S^1) \to \mathcal{T}/\mathcal{K}$ by $\theta: f \mapsto [T_f]$, the class of operators $T_f$. Next we show that $T_f$ is compact if and only if $f\equiv 0$. Suppose $T_f$ is compact. Then for a~continuous~$f$ with Fourier series $ \sum\limits_{n=-\infty}^\infty e_n e^{inx}$ we have \begin{gather*} T_f\big(e^{ikx}\big)=\sum\limits_{n=0}^\infty e_{n-k}e^{inx}. \end{gather*} Thus, the matrix coef\/f\/icients $e_n=(E_{i+n},T_f E_i)$ and since $T_f$ is compact, we must have $(E_{i+n}$, $T_f E_i)\to0$ as $i\to\infty$ for each f\/ixed~$n$. Therefore, $e_n=0$ for all~$n$ and hence $f\equiv 0$. This result means that~$\theta$ is injective. Next we observe that $T_fT_g-T_{fg}$ is a~compact operator for all continuous $f$, $g$. If $f$, $g$ are trigonometric polynomials then a~direct calculation shows that $T_fT_g-T_{fg}$ is a~f\/inite rank operator. The general case then follows by appealing to the Stone--Weierstrass theorem. As a~consequence, the map~$\theta$ above is a~$C^*$-homomorphism. The range of~$\theta$ is dense since it contains (the classes of) polynomials in~$z$ and $z^*$. Then by general $C^*$-algebra theory (see~\cite{Conway} for example)~$\theta$ is an isometry hence the range is closed. This means that $\Ran(\theta)=\mathcal{T}/\mathcal{K}$ and therefore~$\theta$ is a~$*$-isomorphism. \end{proof} Note that from the last theorem we get a~short exact sequence \begin{gather*} 0\rightarrow \mathcal{K}\rightarrow\mathcal{T}\rightarrow C(S^1)\rightarrow 0. \end{gather*} We can compare this to the short exact sequence for the classical disk \begin{gather*} 0\rightarrow C_0({\mathbb D})\rightarrow C({\mathbb D})\rightarrow C(S^1)\rightarrow 0, \end{gather*} where $C_0({\mathbb D})$ are the continuous functions on the disk that vanish on the boundary. \subsection{The quantum annulus} Let $0<r<1$ and consider the annulus \begin{gather*} A_{r}= \left\{\zeta\in{\mathbb C}: r \leq |\zeta| \leq 1\right\}. \end{gather*} The classical uniformization theory of Riemann surfaces implies that every open annulus is biholomorphically equivalent to an annulus of the above form. We can write any holomorphic function $\varphi(\zeta)$ on the interior of $A_{r}$ as the following convergent version of Laurent series \begin{gather*} \varphi(\zeta)=\sum\limits_{n=0}^\infty e_n\zeta^n+\sum\limits_{n=-\infty}^{-1} f_n\left(\frac{\zeta}{r}\right)^n. \end{gather*} We label the basic monomials in the above expansion as \begin{gather*} E_n=\zeta^n, \qquad F_n =\left(\frac{\zeta}{r}\right)^n, \end{gather*} and def\/ine our specially convenient Hilbert space of holomorphic functions on $A_{r}$ to be \begin{gather*} H=\left\{\varphi(\zeta)=\sum\limits_{n=0}^\infty e_nE_n+\sum\limits_{n=-\infty}^{-1}f_nF_n: \|\varphi\|<\infty\right\}, \end{gather*} where \begin{gather*} \|\varphi\|^2=\sum\limits_{n=0}^\infty |e_n|^2+\sum\limits_{n=-\infty}^{-1}|f_n|^2, \end{gather*} so that $\{E_n\}$, $\{F_m\}$ form an orthonormal basis. The operator $z:H \to H$ is def\/ined by the formula $f(\zeta)\mapsto \zeta f(\zeta)$. With respect to the above basis, the operator~$z$ is a~rather special weighted bilateral shift. We have \begin{alignat*}{3} & zE_n=E_{n+1} \qquad && \text{for} \quad n\ge 0, & \\ & zF_n=rF_{n+1} \qquad && \text{for} \quad n\le -2, & \\ & zF_{-1}=rE_0&&& \end{alignat*} and \begin{alignat*}{3} & z^*E_n=E_{n-1} \qquad && \text{for} \quad n\ge 1,& \\ & z^*E_0=rF_{-1},&&& \\ & z^*F_n=rF_{n-1} \qquad && \text{for} \quad n\le -1.& \end{alignat*} In full analogy with the disk case, the operator~$z$ is a~form of a~noncommutative coordinate for what we call quantum annulus. First we look at the spectrum of~$z$. \begin{Theorem} The norm of~$z$ is~$1$. The spectrum of~$z$ is all of~$A_{r}$. \end{Theorem} \begin{proof} The formulas above easily imply that $\|z\|\leq 1$, while the action of~$z$ on $E_n$ shows that it is exactly~1. It is then straightforward to verify that for~$\lambda$ inside $A_{r}$ the following is an eigenvector of $z^*$ corresponding to the eigenvalue~$\lambda$ \begin{gather*} \phi_\lambda=\sum\limits_{n=0}^\infty \lambda^n E_n+\sum\limits_{n=-\infty}^{-1}\left(\frac{\lambda}{r}\right)^nF_n. \end{gather*} Finally, using the techniques described in Lemma~\ref{resolvent_of_z} below, we can prove that the operator $z-\lambda$ is invertible for $|\lambda|<r$. Put together those statements imply that the spectrum of~$z$ is~$A_{r}$. \end{proof} The operators $z^*z$ and $zz^*$ are diagonal. We have \begin{alignat*}{3} & zz^*E_n=E_{n} \qquad && \text{for} \quad n\ge 1,& \\ & zz^*F_n=r^2F_{n} \qquad && \text{for}\quad n\le -1,& \\ & zz^*E_{0}=r^2E_0 &&& \end{alignat*} and \begin{alignat*}{3} & z^*zE_n=E_{n} \qquad && \text{for} \quad n\ge 1,& \\ & z^*zF_n=r^2F_{n} \qquad&& \text{for} \quad n\le -1,& \\ & z^*zE_0=E_{0}.&&& \end{alignat*} Thus the spectrum of those operators is $\sigma(zz^*)=\{1\}\cup\{r^2\}=\sigma(z^*z)$. Also notice that the spectral projections $P_{z^*z}(1)$ and $P_{z^*z}(r^2)$ of $z^*z$ are orthogonal projections onto subspaces of~$H$ generated by $E_n$'s and $F_n$'s, respectively. By the continuous functional calculus applied to $z^*z$, both projections belong to $C^*(z)$, the $C^*$-algebra generated by~$z$. \begin{Remark} The above formulas also imply that the commutator $\frac{z^*z-zz^*}{1-r^2}=\frac{[z^*,z]}{1-r^2}$ is the ortho\-go\-nal projection onto the one-dimensional subspace spanned by $E_0$,hence a~compact operator. \end{Remark} \begin{Theorem} The commutator ideal of $C^*(z)$ is the ideal of compact operators. \end{Theorem} \begin{proof} By the remark above the commutator ideal of $C^*(z)$ is contained in $\mathcal{K}$. Similar to the quantum disk case, the opposite inclusion follows from the easily verif\/iable fact that the rank one operators $f\mapsto\langle f, E_i\rangle E_j$, $f\mapsto\langle f, E_i\rangle F_j$, $f\mapsto\langle f, F_i\rangle E_j$, $f\mapsto\langle f, F_i\rangle F_j$ are in the commutator ideal of $C^*(z)$. \end{proof} \begin{Theorem} The quotient $C^*(z)/\mathcal{K}$ is isomorphic to $C(S^1)\oplus C(S^1)$, where $C(S^1)$ is the space of continuous functions on the unit circle. Thus we have a~short exact sequence \begin{gather*} 0\rightarrow \mathcal{K}\rightarrow C^*(z)\rightarrow C(S^1)\oplus C(S^1)\rightarrow 0. \end{gather*} \end{Theorem} \begin{proof} For details we refer to the proof of Theorem~\ref{quotient_algebra} in the next section. The key step is showing that the inf\/inite-dimensional spectral projections $P_{z^*z}(1)$ and $P_{z^*z}(r^2)$ are in $C^*(z)$. They can be used together with Toeplitz operators on subspaces generated by $E_n$'s and $F_n$'s to construct an isomorphism between $C^*(z)/\mathcal{K}$ and $C(S^1)\oplus C(S^1)$ in a~similar fashion to the Toeplitz algebra case. \end{proof} \section{The quantum pair of pants}\label{section3} Let $0<a<1$, $a+r_2<1$, $r_1+r_2<a$. We def\/ine the (closed) pair of pants as follows \begin{gather*} PP_{(a, r_1,r_2)}= \left\{\zeta\in{\mathbb C}: |\zeta| \leq 1,\, |\zeta|\geq r_1,\, |\zeta-a|\geq r_2\right\}. \end{gather*} It is clear that every open disk with two nonintersecting circular holes is biholomorphically equivalent to the interior of the one of the above pair of pants. There are some technical advantages to having the holes located as above. To a~pair of pants we associate a~convenient Hilbert space of holomorphic functions on it and study the operator of multiplication by~$\zeta$ on that Hilbert space. This is described more precisely in the following subsection. \subsection{Definitions} It follows from~\cite{M} that every holomorphic function on the interior of $PP_{(a, r_1,r_2)}$ can be appro\-xi\-ma\-ted~by rational functions with the only singularities at the centers of the smaller circles in $PP_{(a, r_1,r_2)}$ or at inf\/inity. In fact we can do a~little better. \begin{Proposition} Every holomorphic function $\varphi(\zeta)$ on the interior of $PP_{(a, r_1,r_2)}$ can be written as the following convergent series \begin{gather*} \varphi(\zeta)=\sum\limits_{n=0}^\infty e_n\zeta^n+\sum\limits_{n=-\infty}^{-1} f_n\left(\frac{\zeta}{r_1}\right)^n + \sum\limits_{n=-\infty}^{-1} g_n\left(\frac{\zeta-a}{r_2}\right)^n. \end{gather*} \end{Proposition} \begin{proof} In~\cite{Ahlfors}, it was shown that if $\varphi(\zeta)$ is holomorphic on an annulus $\{\zeta\in{\mathbb C}: R_1 < |\zeta-c|<R_2\}$, then $\varphi(\zeta)=\varphi_1(\zeta)+\varphi_2(\zeta)$ where $\varphi_1(\zeta)$ is holomorphic on $|\zeta-c|>R_1$ and $\varphi_2(\zeta)$ is holomorphic on $|\zeta-c|<R_2$. We apply this theorem twice. Let~$\varphi$ be a~holomorphic function on the open pair of pants. Consider an annulus $A=\{\zeta\in{\mathbb C}: |\zeta-a|>r_2, |\zeta-c|<r\}$ around~$a$ with outer radius~$r$ and inner radius~$r_2$ that does not intersect the hole around the origin with radius~$r_1$ and let~$D$ be the disk with center~$a$ and radius~$r$. Then $\varphi|_{A}$ is a~holomorphic function and so from~\cite{Ahlfors}, $\varphi|_{A}=\varphi_1+\varphi_2$ with $\varphi_1$ holomorphic outside the hole centered at~$a$ with radius~$r_2$, and~$\varphi_2$ holomorphic on~$D$. Consequently~$\varphi_1$ has the following convergent series representation \begin{gather*} \varphi_1=\sum\limits_{n=-\infty}^{-1} g_n\left(\frac{\zeta-a}{r_2}\right)^n. \end{gather*} Next consider the function $\varphi-\varphi_1$. This function is holomorphic on $PP_{(a, r_1,r_2)}$ and, because $\varphi-\varphi_1=\varphi_2$ on A, it extends to a~holomorphic function on~$D$. This means that $\varphi-\varphi_1$ is holomorphic on the annulus $\{\zeta\in{\mathbb C}: |\zeta|>r_1, |\zeta-c|<1\}$, and so by using~\cite{Ahlfors} again, we have $\varphi-\varphi_1=\varphi_3+\varphi_4$ with $\varphi_3$ holomorphic in the unit disk ${\mathbb D}$ and $\varphi_4$ holomorphic on $\{\zeta\in{\mathbb C}: |\zeta|>r_1 \}$. Thus $\varphi_3$ and $\varphi_4$ have the following convergent series representation \begin{gather*} \varphi_3=\sum\limits_{n=0}^\infty e_n \zeta^n \qquad \text{and} \qquad \varphi_4=\sum\limits_{n=-\infty}^{-1}f_n\left(\frac{\zeta}{r_1}\right)^n. \end{gather*} Combining these three series representations gives the desired result. \end{proof} Similar to the annulus case we set \begin{gather*} E_n=\zeta^n, \qquad F_n =\left(\frac{\zeta}{r_1}\right)^n, \qquad \text{and} \qquad G_n=\left(\frac{\zeta-a}{r_2}\right)^n. \end{gather*} The Hilbert space~$H$ that we will use is def\/ined as \begin{gather} \label{Hilbert_space} H=\left\{\varphi(\zeta)=\sum\limits_{n=0}^\infty e_nE_n+\sum\limits_{n=-\infty}^{-1}(f_nF_n+g_nG_n): \|\varphi\|<\infty\right\}, \end{gather} where \begin{gather*} \|\varphi\|^2=\sum\limits_{n=0}^\infty |e_n|^2+\sum\limits_{n=-\infty}^{-1}\left(|f_n|^2+|g_n|^2\right). \end{gather*} The advantage of working with the above Hilbert space of holomorphic functions on $PP_{(a, r_1,r_2)}$ is that there is a~distinguished orthonormal basis in it, namely the basis consisting of $\{E_n\}$, $\{F_m\}$, $\{G_k\}$. The object of study in this section is the operator $z:H \to H$ given by $z\varphi(\zeta)=M_\zeta \varphi(\zeta) = \zeta \varphi(\zeta)$, i.e., the multiplication operator by~$\zeta$. Straightforward calculations yields the following formulas. \begin{Lemma} The operators~$z$ and $z^*$ act on the basis elements in the following way \begin{alignat*}{3} & zE_n=E_{n+1} \qquad && \text{for} \quad n\ge 0,& \\ & zF_n=r_1F_{n+1} \qquad && \text{for} \quad n\le -2, & \\ & zF_{-1}=r_1E_0,&&& \\ & zG_n=r_2G_{n+1}+aG_n \qquad && \text{for} \quad n\le -2, & \\ & zG_{-1}=r_2E_0+aG_{-1}&&& \end{alignat*} and \begin{alignat*}{3} & z^*E_n=E_{n-1} \qquad && \text{for} \quad n\ge 1,& \\ & z^*E_0=r_1F_{-1}+r_2G_{-1}, &&& \\ & z^*F_n=r_1F_{n-1} \qquad && \text{for} \quad n\le -1,& \\ & z^*G_n=r_2G_{n-1}+aG_n \qquad && \text{for} \quad n\le -1. & \end{alignat*} \end{Lemma} \begin{Lemma} \label{z_z_star_coefficients} The operators~$z$ and $z^*$ shift the coefficients of $\varphi(\zeta)$ in the series decomposition defined in equation~\eqref{Hilbert_space} in the following way \begin{gather*} z\varphi=\sum\limits_{n=0}^\infty\tilde{e}_nE_n+\sum\limits_{n=-\infty}^{-1}\!\! \big(\tilde{f}_nF_n+\tilde{g}_nG_n\big) \qquad \text{and} \qquad z^*\varphi=\sum\limits_{n=0}^\infty e'_nE_n+\sum\limits_{n=-\infty}^{-1}\!\! (f'_nF_n+g'_nG_n), \end{gather*} where \begin{alignat*}{3} & \tilde{e}_n=e_{n-1} \qquad && \text{for} \quad n\ge 1, & \\ & \tilde{e}_0=r_1f_{-1}+r_2g_{-1}, &&& \\ & \tilde{f}_n=r_1f_{n-1} \qquad && \text{for} \quad n\le -1,& \\ & \tilde{g}_n=r_2g_{n-1}+ag_n \qquad && \text{for} \quad n\le -1 & \end{alignat*} and \begin{alignat*}{3} & e'_n=e_{n+1} \qquad && \text{for} \quad n\ge 0, & \\ & f'_n=r_1f_{n+1} \qquad && \text{for} \quad n\le -2, & \\ & f'_{-1}=r_1e_0, &&& \\ & g'_n=r_2g_{n+1}+ag_n \qquad && \text{for} \quad n\le -2,& \\ & g'_{-1}=r_2e_0+ag_{-1}. &&& \end{alignat*} \end{Lemma} We can now def\/ine the quantum pair of pants. \begin{Definition} The quantum pair of pants, denoted $QPP_{(a, r_1,r_2)}$, is def\/ined to be the $C^*$-algebra generated by the operator~$z$, i.e., $QPP_{(a, r_1,r_2)}=C^*(z)$. \end{Definition} \subsection[The spectrum of~$z$]{The spectrum of~$\boldsymbol{z}$} In this subsection we study the spectrum of~$z$, starting with a~calculation of the norm of~$z$. \begin{Proposition} \label{norm_of_z} With the above notation, we have: $\|z\|=1$. \end{Proposition} \begin{proof} Using the series representation of $\varphi(\zeta)$ in formula \eqref{Hilbert_space} above, and the coef\/f\/icients of Lemma~\ref{z_z_star_coefficients} we compute $\|z\varphi\|^2$: \begin{gather*} \|z\varphi\|^2=\sum\limits_{n=1}^\infty |e_{n-1}|^2+|r_1f_{-1}+r_2g_{-1}|^2+r_1^2\sum\limits_{n=-\infty}^{-1} |f_{n-1}|^2+\sum\limits_{n=-\infty}^{-1}|r_2g_{n-1}+ag_n|^2. \end{gather*} Using the triangle inequality and the fact that $a> r_1$ we obtain \begin{gather*} \|z\varphi\|^2 \le \sum\limits_{n=0}^\infty |e_n|^2+\big(a|f_{-1}|+r_2|g_{-1}|\big)^2+r_1^2 \sum\limits_{n=-\infty}^{-1}|f_{n-1}|^2+\sum\limits_{n=-\infty}^{-1}\big(r_2|g_{n-1}|+a|g_n|\big)^2. \end{gather*} Notice that by denoting $g_0:= f_{-1}$ we can write \begin{gather*} \big(a|f_{-1}|+r_2|g_{-1}|\big)^2+\sum\limits_{n=-\infty}^{-1}\big(r_2|g_{n-1}|+a|g_n|\big)^2= \sum\limits_{n=-\infty}^{0}\big(r_2|g_{n-1}|+a|g_n|\big)^2 \\ \qquad{} =r_2^2 \sum\limits_{n=-\infty}^0 |g_{n-1}|^2+ a^2\sum\limits_{n=-\infty}^0 |g_n|^2+2ar_2 \sum\limits_{n=-\infty}^0 |g_{n-1}||g_n|. \end{gather*} The Cauchy--Schwartz inequality implies \begin{gather*} \sum\limits_{n=-\infty}^{0}\big(r_2|g_{n-1}|+a|g_n|\big)^2 \leq r_2^2 \sum\limits_{n=-\infty}^0 |g_{n-1}|^2+ a^2\sum\limits_{n=-\infty}^0 |g_n|^2 \\ \qquad \phantom{\leq} {} +2ar_2 \left(\sum\limits_{n=-\infty}^0 |g_{n-1}|^2\right)^{1/2}\left(\sum\limits_{n=-\infty}^0 |g_n|^2\right)^{1/2} \\ \qquad \leq \big(r_2^2+a^2+2ar_2\big) \sum\limits_{n=-\infty}^0 |g_n|^2 =(r_2+a)^2 \left(|f_{-1}|^2+\sum\limits_{n=-\infty}^{-1} |g_n|^2\right). \end{gather*} Using the fact that $r_1, r_2+a <1$ in the above computations we see that \begin{gather*} \|z\varphi\|^2 \le \sum\limits_{n=0}^\infty |e_n|^2+r_1^2\sum\limits_{n=\infty}^{-1}|f_{n-1}|^2+(r_2+a)^2 |f_{-1}|^2+(r_2+a)^2\sum\limits_{n=-\infty}^{-1}|g_n|^2 \\ \phantom{\|z\varphi\|^2}{} \leq \sum\limits_{n=0}^\infty |e_n|^2+\sum\limits_{n=\infty}^{-1}|f_n|^2+\sum\limits_{n=-\infty}^{-1}|g_n|^2 = \|\varphi\|^2, \end{gather*} showing that $\|z\|\le 1$. On the other hand, $\|zE_1\| =\| E_2\|=\|E_1\|$. Thus $\|z\|=1$. \end{proof} Next we compute the spectrum of~$z$. In estimating the norms of resolvents of~$z$ we use the following well known result. \begin{Lemma}[Schur--Young inequality] \label{schuryounginq} Let $T: L^2(Y) \longrightarrow L^2(X)$ be an integral operator \begin{gather*} Tf(x)=\int K(x,y)f(y)dy. \end{gather*} Then one has \begin{gather*} \|T\|^2 \le \left(\sup_{x\in X} \int_Y |K(x,y)|dy\right)\left(\sup_{y\in Y} \int_X |K(x,y)|dx\right). \end{gather*} \end{Lemma} The details of the lemma and its proof can be found in~\cite{HS}. \begin{Lemma} \label{resolvent_of_z} The operator $z-\lambda$ has a~bounded inverse for $|\lambda|< r_1$, $|\lambda-a|< r_2$, and $|\lambda|>1$. \end{Lemma} \begin{proof} Let \begin{gather*} \varphi(\zeta)=\sum\limits_{n=0}^\infty e_n E_n+\sum\limits_{n=-\infty}^{-1}(f_nF_n+g_nG_n) \end{gather*} and \begin{gather*} \tilde{\varphi}(\zeta)=\sum\limits_{n=0}^\infty \tilde{e}_n E_n+\sum\limits_{n=-\infty}^{-1}\big(\tilde{f}_nF_n+\tilde{g}_nG_n\big). \end{gather*} Consider the equation $(z-\lambda)\varphi(\zeta)=\tilde{\varphi}(\zeta)$. Using the above decompositions and Lemma~\ref{z_z_star_coefficients} we obtain the following system of equations \begin{alignat}{3} & r_1f_{-1}+r_2g_{-1}-\lambda e_0=\tilde{e}_0,\qquad &&&\nonumber \\ & e_{n-1}-\lambda e_n=\tilde{e}_n \quad &&\text{for} \quad n\ge1, &\nonumber \\ & r_1f_{n-1}-\lambda f_n=\tilde{f}_n \quad &&\text{for} \quad n\le-1,&\nonumber \\ & r_2g_{n-1}+ag_n-\lambda g_n=\tilde{g}_n \quad &&\text{for} \quad n\le-1. & \label{resolvent_system} \end{alignat} By Proposition~\ref{norm_of_z}, $\|z\|=1$ and if $|\lambda|>1=\|z\|$ then by general functional analysis we know that $(z-\lambda)^{-1}$ is a~bounded, invertible operator. Next we consider three cases: the f\/irst case is for $0<|\lambda|<r_1$, the second case is for $|\lambda-a|<r_2$, and the last case is for $\lambda=0$. If $0<|\lambda|<r_1<1$, then $|\lambda-a|>r_2$. We can solve the system of equations~\eqref{resolvent_system} recursively. Rewriting the last equation and multiplying by $((\lambda-a)/r_2)^{n-1}$ yields \begin{gather*} \left(\frac{\lambda-a}{r_2}\right)^{n-1}g_{n-1}-\left(\frac{\lambda-a}{r_2}\right)^ng_n =\left(\frac{\lambda-a}{r_2}\right)^{n-1}\frac{1}{r_2}\tilde{g}_n. \end{gather*} Letting $h_n=((\lambda-a)/r_2)^ng_n$, we get \begin{gather*} h_{n-1}-h_n=\left(\frac{\lambda-a}{r_2}\right)^{n-1}\frac{1}{r_2}\tilde{g}_n. \end{gather*} The requirement for a~square summable solution forces $h_n = -\sum\limits_{j=-\infty}^n((\lambda-a)/r_2)^{j-1}\tilde{g}_j/r_2$ and hence for $n\le-1$ we obtain \begin{gather*} g_n=-\frac{1}{r_2}\sum\limits_{j=-\infty}^n\left(\frac{\lambda-a}{r_2}\right)^{j-n-1}\tilde{g}_j. \end{gather*} Similar calculations show that \begin{gather*} e_n=\sum\limits_{j=n+1}^\infty \lambda^{j-n-1}\tilde{e}_j \qquad \text{and} \qquad f_n=\left(\frac{\lambda}{r_1}\right)^{-n-1}f_{-1}+\frac{1}{r_1}\sum\limits_{j=n+1}^{-1}\left(\frac{\lambda}{r_1}\right)^{j-n-1}\tilde{f}_j \end{gather*} for $n\ge0$ and $n\le-2$ respectively. These formulas along with the f\/irst equation in system~\eqref{resolvent_system} give \begin{gather*} f_{-1}=\frac{1}{r_1}\left(\sum\limits_{j=0}^\infty\lambda^j\tilde{e}_j+\sum\limits_{j=-\infty}^{-1}\left(\frac{\lambda - a}{r_2}\right)^j\tilde{g}_j\right). \end{gather*} We introduce some notation; f\/irst notice that we have a~natural decomposition, $H \!\cong\! \ell^2({\mathbb Z}_{\ge0}) \oplus \ell^2({\mathbb Z}_{<0}) \oplus \ell^2({\mathbb Z}_{<0})$ given in the following way: for $\varphi\in H$ write $\varphi=e+f+g$ where $e=\sum\limits_{n\ge0}e_nE_n$, $f= \sum\limits_{n\le-1}f_nF_n$, and $g=\sum\limits_{n\le-1}g_nG_n$. Using this notation we see that $\|\varphi\|^2=\|e\|^2+\|f\|^2+\|g\|^2$. Def\/ine the characteristic $\chi(t)=1$ for $0\le t\le1$ and zero otherwise, then we can def\/ine seven dif\/ferent integral operators \begin{gather} T_1e=\sum\limits_{n=0}^\infty\sum\limits_{j=0}^\infty \lambda^{j-n-1}\chi\left(\frac{n+1}{j}\right)e_jE_n: \ \ell^2({\mathbb Z}_{\ge0})\to\ell^2({\mathbb Z}_{\ge0}), \nonumber \\ T_2(e,g)=\sum\limits_{n=-\infty}^{-1}\frac{1}{r_1}\left(\frac{\lambda}{r_1}\right)^{-n-1}\left(\sum\limits_{j=0}^\infty\lambda^je_j + \sum\limits_{j=-\infty}^{-1}\left(\frac{\lambda-a}{r_2}\right)^jg_j\right)F_n : \nonumber \\ \hphantom{T_2(e,g)=}{} \ell^2({\mathbb Z}_{\ge0})\oplus\ell^2({\mathbb Z}_{<0})\to\ell^2({\mathbb Z}_{<0}) \nonumber \\ T_3f = \sum\limits_{n=-\infty}^{-1}\frac{1}{r_1}\sum\limits_{j=-\infty}^{-1}\left(\frac{\lambda}{r_1}\right)^{j-n-1}\chi\left(\frac{n+1}{j}\right)f_jF_n: \ \ell^2({\mathbb Z}_{<0})\to\ell^2({\mathbb Z}_{<0}), \nonumber \\ \label{int_ops_resolvent_z} T_4f = \sum\limits_{n=-\infty}^{-1}\frac{1}{r_1}\sum\limits_{j=-\infty}^{-1}\left(\frac{\lambda}{r_1}\right)^{j-n-1}\chi\left(\frac{j}{n}\right)f_jF_n: \ \ell^2({\mathbb Z}_{<0})\to\ell^2({\mathbb Z}_{<0}), \\ T_5(e,f) =\sum\limits_{n=-\infty}^{-1}\frac{1}{r_2}\left(\frac{\lambda-a}{r_2}\right)^{-n-1} \left(\sum\limits_{j=0}^\infty\lambda^je_j+\sum\limits_{j=-\infty}^{-1}\left(\frac{\lambda}{r_1}\right)^jf_j\right)G_n: \nonumber \\ \hphantom{T_5(e,f) =}{} \ell^2({\mathbb Z}_{\ge0})\oplus\ell^2({\mathbb Z}_{<0})\to\ell^2({\mathbb Z}_{<0}), \nonumber \\ T_6g=\sum\limits_{n=-\infty}^{-1}\frac{1}{r_2}\sum\limits_{j=-\infty}^{-1}\left(\frac{\lambda-a}{r_2}\right)^{j-n-1} \chi\left(\frac{n+1}{j}\right)g_j G_n: \ \ell^2({\mathbb Z}_{<0})\to\ell^2({\mathbb Z}_{<0}), \nonumber \\ T_7g=\sum\limits_{n=-\infty}^{-1}\frac{1}{r_2}\sum\limits_{j=-\infty}^{-1}\left(\frac{\lambda-a}{r_2}\right)^{j-n-1} \chi\left(\frac{j}{n}\right)g_j G_n: \ \ell^2({\mathbb Z}_{<0})\to\ell^2({\mathbb Z}_{<0}). \nonumber \end{gather} The operators from formula~\eqref{int_ops_resolvent_z} can be used to represent $(z-\lambda)^{-1}\tilde{\varphi}$, for $\tilde{\varphi}=\tilde{e}+\tilde{f}+\tilde{g}$ where $\tilde{e}=\sum\limits_{n\ge0}\tilde{e}_nE_n$, $\tilde{f}=\sum\limits_{n\le -1}\tilde{f}_nF_n$ and $\tilde{g}=\sum\limits_{n\le -1}\tilde{g}_nG_n$, in the following way \begin{gather*} (z-\lambda)^{-1}\tilde{\varphi}=T_1\tilde{e}+T_2(\tilde{e},\tilde{g})+T_3\tilde{f}-T_7\tilde{g}. \end{gather*} Next we estimate the norm of $(z-\lambda)^{-1}$. We use Lemma~\ref{schuryounginq} to estimate the norms of the operators $T_1$, $T_3$ and $T_7$ and we directly estimate $\|T_2\tilde{f}\|$. The f\/irst estimate is \begin{gather*} \|T_1\|^2 \leq \left(\sup_{n \geq 0} |\lambda|^{-n-1} \sum\limits_{j=n+1}^{\infty}|\lambda|^j \right)\left(\sup_{j \geq 1} |\lambda|^{j-1} \sum\limits_{n=0}^{j-1}|\lambda|^{-n} \right) \\ \phantom{\|T_1\|^2}{} =\frac 1{\left(1-|\lambda|\right)}\left(\sup_{j \geq 1} \frac{1-|\lambda|^j}{1-|\lambda|} \right)=\frac1{\left(1-|\lambda|\right)^2}, \end{gather*} where we have used the fact that $|\lambda|<1$. Similarly, we have \begin{gather*} \|T_3\|^2 \leq \frac 1{r_1^2}\left(\sup_{n \leq -2} \left(\frac{|\lambda|}{r_1}\right)^{-n-1} \sum\limits_{j=n+1}^{-1}\left(\frac{|\lambda|}{r_1}\right)^j \right) \left(\sup_{j \leq -1} \left(\frac{|\lambda|}{r_1}\right)^{j-1} \sum\limits_{n=-\infty}^{j-1}\left(\frac{|\lambda|}{r_1}\right)^{-n} \right) \\ \phantom{\|T_3\|^2}{} \leq \frac 1{r_1^2 \big(1- \frac{|\lambda|}{r_1}\big)^2} \left(\sup_{n\leq -2} 1-\left(\frac{|\lambda|}{r_1}\right)^{-n-1}\right). \end{gather*} Since $\frac{|\lambda|}{r_1} <1$, it follows that $\|T_3\|^2 \leq \frac{1}{r_1^2\big(1- \frac{|\lambda|}{r_1}\big)^2}$. Next, \begin{gather*} \|T_7\|^2 \leq \frac 1{r_2^2} \left(\sup_{n \leq -1} \left|\frac{\lambda -a}{r_2} \right|^{-n-1} \sum\limits_{j=-\infty}^n \left|\frac{\lambda -a}{r_2} \right|^j \right) \left(\sup_{j \leq -1} \left|\frac{\lambda -a}{r_2} \right|^{j-1} \sum\limits_{n=j}^{-1} \left|\frac{\lambda -a}{r_2} \right|^{-n} \right) \\ \phantom{\|T_7\|^2}{} \leq \frac 1{r_2^2} \frac{(|\lambda -a|/ r_2)^{-2}}{\big(1- (|\lambda -a|/ r_2)^{-1}\big)^2} \left(\sup_{j \leq -1} 1- \left(\frac{|\lambda -a|}{r_2}\right)^j\right). \end{gather*} Because $\frac{|\lambda -a|}{r_2}>1$, we have \begin{gather*} \|T_7\|^2 \leq \frac 1{r_2^2} \frac{(|\lambda -a|/ r_2)^{-2}}{\big(1- (|\lambda -a|/ r_2)^{-1}\big)^2}=\frac {1}{r_1^2 \big(\frac{|\lambda -a|}{r_2} -1\big)^2}. \end{gather*} The operator $T_2$ is a~rank one operator and the norm $T_2\tilde{f}$ can be estimated directly, using the Cauchy--Schwartz inequality \begin{gather*} \|T_2(\tilde{e},\tilde{g})\|^2= \frac{1}{r_1^2}\sum\limits_{n=-\infty}^{-1}\left(\frac{|\lambda|}{r_1}\right)^{-2n-2}\left|\sum\limits_{j=0}^\infty\lambda^j\tilde{e}_j + \sum\limits_{j=-\infty}^{-1}\left(\frac{\lambda-a}{r_2}\right)^j\tilde{g}_j\right|^2 \\ \phantom{\|T_2(\tilde{e},\tilde{g})\|^2}{} \le\frac{1}{r_1^2(1-(|\lambda|/r_1)^2)}\left(\frac{1}{1-|\lambda|^2}+ \frac{1}{(|\lambda-a|/r_2)^2-1}\right) \left(\|\tilde{e}\|^2+\|\tilde{g}\|^2\right). \end{gather*} This shows that $(z-\lambda)^{-1}$ is bounded for $0<|\lambda|<r_1$. The second case is $|\lambda-a|<r_2$. This implies that $r_1<|\lambda|<1$. Under these constraints we solve system~\eqref{resolvent_system} using the same methods as those for the f\/irst case to obtain \begin{gather*} e_n =\sum\limits_{j=n+1}^\infty \lambda^{j-n-1}\tilde{e}_j \qquad\text{for}\quad n\ge0, \\ f_n =-\frac{1}{r_1}\sum\limits_{j=-\infty}^n \left(\frac{\lambda}{r_1}\right)^{j-n-1}\tilde{f}_j \qquad\text{for}\quad n\le-1, \\ g_n =-\left(\frac{\lambda-a}{r_2}\right)^{-n-1}g_{-1}+\frac{1}{r_2}\sum\limits_{j=n+1}^{-1}\left(\frac{\lambda-a}{r_2}\right)^{j-n-1}\tilde{g}_j \qquad \text{for} \quad n\le-2. \end{gather*} Then the f\/irst equation of system~\eqref{resolvent_system} gives \begin{gather*} g_{-1}=\frac{1}{r_2}\left(\sum\limits_{j=0}^\infty\lambda^j\tilde{e}_j+\sum\limits_{j=-\infty}^{-1}\left(\frac{\lambda}{r_1}\right)^j\tilde{f}_j\right). \end{gather*} Similar to the f\/irst case we can express $(z-\lambda)^{-1}$ using the operators def\/ined in formula~\eqref{int_ops_resolvent_z} to get \begin{gather*} (z-\lambda)^{-1}\tilde{\varphi}=T_1\tilde{e}-T_4\tilde{f}+T_5(\tilde{e},\tilde{f})+T_6\tilde{g}. \end{gather*} We omit the repetitive details of estimates of $T_4$, $T_5$, and $T_6$ norms. They imply that $(z-\lambda)^{-1}$ is bounded for $|\lambda -a|<r_2$. The last case is when $\lambda=0$. Solving system~\eqref{resolvent_system} we obtain \begin{gather*} e_n =\tilde{e}_{n+1} \qquad\text{for}\quad n\ge0, \\ f_n =\frac{1}{r_1}\tilde{f}_{n+1} \qquad\text{for}\quad n\le-2, \\ g_n =-\frac{1}{r_2}\sum\limits_{j=-\infty}^n\left(-\frac{a}{r_2}\right)^{j-n-1}\tilde{g}_j \qquad\text{for}\quad n\le-1. \end{gather*} Using the f\/irst equation of system~\eqref{resolvent_system} we compute $f_{-1}$ \begin{gather*} f_{-1}=\frac{1}{r_1}\left(\tilde{e}_0+\sum\limits_{j=-\infty}^{-1}\left(-\frac{a}{r_2}\right)^j\tilde{g}_j\right). \end{gather*} As before the norm estimates hinge on convergent geometric series. This completes the proof. \end{proof} \begin{Theorem The spectrum of~$z$ is the regular pair of pants, i.e., $\sigma(z)=PP_{(a, r_1,r_2)}$. \end{Theorem} \begin{proof} By Proposition~\ref{norm_of_z}, $\sigma(z)\subset {\mathbb D}$. Let \begin{gather*} \varphi_\lambda(\zeta)=\sum\limits_{n=0}^\infty \lambda^nE_n+\sum\limits_{n=-\infty}^{-1} \left(\frac{\lambda}{r_1}\right)^nF_n+\sum\limits_{n=-\infty}^{-1}\left(\frac{\lambda-a}{r_2}\right)^nG_n. \end{gather*} It is easy to see that for any~$\lambda$ in the interior of $PP_{(a, r_1,r_2)}$, $\varphi_\lambda\in H$ and that~$\lambda$ is an eigenvalue with $\varphi_\lambda$ as the associated eigenfunction for $z^*$. Therefore, $PP_{(a, r_1,r_2)}\subset\sigma(z)$. From Lemma~\ref{resolvent_of_z} the operator $z-\lambda$ has a~bounded inverse whenever $|\lambda|< r_1$ or $|\lambda-a|<r_2$. Hence the resolvent set is contained in the holes within the unit disk or outside the unit disk, and so $\sigma(z)\subset PP_{(a, r_1,r_2)}$. \end{proof} In view of the above theorem we can think of the operator~$z$ as a~form of a~noncommutative complex coordinate for what we call quantum pair of pants. \subsection[Structure of $C^*(z)$]{Structure of $\boldsymbol{C^*(z)}$} Next we study the commutator ideal of $C^*(z)$. A~straightforward computation gives the following formulas. \begin{Lemma} \label{commutator_calculation} The commutator of $z^*$ and~$z$ act on the basis elements in the following way: $[z^*,z]E_n=0$ for $n\ge 1$, $[z^*,z]F_n=0$ for $n\le -2$, $[z^*,z]G_n =0$ for $n\le -2$. Moreover on the initial elements we get: \begin{gather*} [z^*,z]E_0=\big(1-r_1^2-r_2^2\big)E_0-ar_2G_{-1}, \\ [z^*,z]F_{-1} =r_1r_2G_{-1}, \\ [z^*,z]G_{-1} =r_1r_2F_{-1}-ar_2E_0. \end{gather*} \end{Lemma} Let $\mathcal{I}$ be the ideal generated by $[z^*,z]$. It is easy to see that $\mathcal{I}$ is in fact the commutator ideal of $C^*(z)$ because that algebra is singly generated. \begin{Theorem The commutator ideal $\mathcal{I}$ of $C^*(z)$ is the $C^*$-algebra $\mathcal{K}$ of compact operators in~$H$. \end{Theorem} \begin{proof} From Lemma~\ref{commutator_calculation} it is clear that the commutator $[z^*, z]$ is f\/inite-rank and hence compact. Thus, $\mathcal I \subset \mathcal K$. On the other hand, to show that $\mathcal K \subset \mathcal I$ we will use the following step by step method building up to the conclusion that a~large collection of rank one operators belong to $\mathcal I$ and that the compact operators are exactly the norm limit of those. {\bf Step 1.} First we show that $P=$ orthogonal projection onto $\operatorname{span}\{E_0, F_{-1}, G_{-1}\}$ belongs to the commutator ideal. Notice that the (self-adjoint) operator $[z^*{,} z]$ acting on $\operatorname{span}\{E_0, F_{{-}1}, G_{{-}1}\}$ has the following matrix representation in the basis $\{E_0, F_{-1}, G_{-1}\}$: \begin{gather*} A= \left( \begin{matrix} 1-r_1^2-r_2^2 & 0 & -ar_2 \\ 0 & 0 & r_1r_2 \\ -ar_2 & r_1r_2 & 0 \end{matrix} \right). \end{gather*} This matrix has rank equal to $3$ and the following characteristic polynomial \begin{gather*} p_A(\lambda)=\lambda^3-\big(1-r_1^2-r_2^2\big)\lambda^2-\big(a^2r_2^2+r_1^2r_2^2\big)\lambda+\big(1-r_1^2-r_2^2\big)r_1^2r_2^2. \end{gather*} Since $0 < r_1$, $r_2 < 1$ it is clear that zero is not an eigenvalue of~$A$. If $\lambda_i$, $i=1,2,3$ are the roots of $p_A(\lambda)$ then by functional calculus there exists a~continuous function $f: \mathbb R \to \mathbb R$ such that $f(0)=0$, $f(\lambda_i)=1$ so that $f([z^*,z])=P$. Consequently $P \in \mathcal I \subset C^*(z)$. {\bf Step 2.} The next step is showing that $P_{E_1}=$ orthogonal projection onto span of $\{E_1\}$ belongs to $\mathcal I$. We f\/irst observe that the operator $zPz^*$ acts on the basis elements in the following way \begin{gather*} zpz^* B= \begin{cases} \big(r_1^2+r_2^2\big)E_0+ar_2G_{-1} & \text{if}\quad B=E_0, \\ E_1 & \text{if}\quad B=E_1, \\ ar_2E_0+a^2 G_{-1} & \text{if}\quad B=G_{-1}, \\ 0 & \text{otherwise}. \end{cases} \end{gather*} Thus, the operator $zPz^*$ on $\operatorname{span}\{E_0, G_{-1}\}$ is self-adjoint and has the following matrix representation in the basis $\{E_0, G_{-1}\}$: \begin{gather*} C=\left( \begin{matrix} r_1^2+r_2^2 & ar_2 \\ ar_2 & a^2 \end{matrix} \right). \end{gather*} The characteristic polynomial for~$C$ is $p_C(\lambda)=\lambda^2-(r_1^2+r_2^2+a^2)\lambda+a^2r_1^2$. First we need to show that $\lambda=0,1$ are not roots of $p_C(\lambda)$. Clearly $p_C(0) \neq 0$. Suppose $p_C(1)=0$. Then solving the equation for $r_2$ we obtain, $r_2^2=(1-a^2)(1-r_1^2)$. Since $r_2 < 1-a$ and $r_2 < 1-r_1$ we see that \begin{gather*} \big(1-a^2\big)\big(1-r_1^2\big) < (1-a)(1-r_1), \end{gather*} implying $(1+a)(1+r_1)<1$ which is clearly a~contradiction since $a, r_1 >0$. Thus, $p_C(1)\neq 0$. Now we look at the discriminant~$\Delta$ of $p_C(\lambda)$: \begin{gather*} \Delta=\big(r_1^2+r_2^2+a^2\big)^2-4a^2r_1^2. \end{gather*} If $\Delta=0$ this would imply that $a=r_1$ and $r_2=0$, which is a~contradiction. Hence~$C$ has two distinct eigenvalues $\lambda_1$ and $\lambda_2$. Thus, once again by functional calculus there exists a~continuous real valued function~$f$ such that $f(0)= f(\lambda_1)=f(\lambda_2)=0$ and $f(1)=1$. Consequently, applying~$f$ to $zPz^*$ we get that, $f(zPz^*)= p_{E_1} \in \mathcal I$. {\bf Step 3.} By similar functional calculus argument as above we also see that $P_{E_0, G_{-1}}$, the orthogonal projection onto $\operatorname{span}\{E_0, G_{-1}\}$, belongs to $\mathcal I$. Consequently, if $P_{F_{-1}}=$ orthogonal projection onto span of $\{F_{-1}\}$ then clearly, $P_{F_{-1}}= P- P_{E_0, G_{-1}} \in \mathcal I$. {\bf Step 4.} We will show that $P_{E_n}= $ orthogonal projection onto $\operatorname{span}\{E_n\}$ belongs to $\mathcal I$ for $n=1,2, \dots$. To this end, we compute the action of $z^{n-1} P_{E_1} (z^*)^{n-1}$ on the basis elements \begin{gather*} z^{n-1} P_{E_1} (z^*)^{n-1} B= \begin{cases} E_n & \text{if}\quad B=E_n, \\ 0 & \text{otherwise.} \end{cases} \end{gather*} Therefore, $z^{n-1} P_{E_1} (z^*)^{n-1}= P_{E_n} \in \mathcal I$ for $n \geq 1$. {\bf Step 5.} Now consider the action of $(z^*)^n P_{F_{-1}} z^n$, $n \geq 1$ on basis elements \begin{gather*} (z^*)^n P_{F_{-1}} z^n B= \begin{cases} r_1^{2n}F_{-n-1} & \text{if}\quad B=F_{-n-1}, \\ 0 & \text{otherwise.} \end{cases} \end{gather*} Thus, $r_1^{2n}(z^*)^{-n} P_{F_{-1}} z^{-n}= P_{F_n}$, the projection onto $F_n$, for $n \leq -1$ and hence the projec\-tions~$P_{F_n}$ belong to $\mathcal I$. {\bf Step 6.} Since $z^* P_{E_1}z E_0=E_0$ and zero elsewhere, it is clear that $z^* P_{E_1}z =P_{E_0} \in \mathcal I$. Hence $P_{G_{-1}}= P- P_{E_0}- P_{F_{-1}}$ also belongs to~$\mathcal I$. {\bf Step 7.} It remains to show that for $n \leq -2$ the orthogonal projection $P_{G_n}$ onto $G_n$ belongs to~$\mathcal I$. We consider the action of $z^* P_{G_{-1}}z$ on basis elements \begin{gather*} z^* P_{G_{-1}}z B= \begin{cases} a^2 G_{-1}+ar_2G_{-2} & \text{if}\quad B=G_{-1}, \\ ar_2 G_{-1}+r_2^2G_{-2} & \text{if}\quad B= G_{-2}, \\ 0 & \text{otherwise}. \end{cases} \end{gather*} Thus it has the following matrix representation relative to the basis $\{G_{-1}, G_{-2}\}$: \begin{gather*} D=\left( \begin{matrix} a^2 & ar_2 \\ ar_2 & r_2^2 \end{matrix} \right). \end{gather*} The matrix~$D$ has eigenvalues $\lambda_1=0$ and $\lambda_2= a^2+r_2^2$ with $v= aG_{-1}+ r_2 G_{-2}$ being the eigenvector corresponding to $\lambda_2$. Thus $P_{v}$, the one-dimensional orthogonal projection onto~$v/\|v\|$, belongs to~$\mathcal I$. Since $P_{G_{-1}}$ and $P_{v}$ are not mutually orthogonal, simple matrix algebra shows that the set $\{I, P_{G_{-1}}, P_{v}, P_{G_{-1}}P_{v}\}$, where~$I$ is the $2 \times 2$ identity matrix, generates the set of all $2 \times 2 $ matrices. Consequently, $P_{G_{-2}}$ can be written as a~linear combination of these four projections, making it clear that $P_{G_{-2}} \in \mathcal I$. Finally we use induction on~$n$ and follow a~similar argument as above to show that $P_{G_n} \in \mathcal I$ for $n=-2,-3, -4, \dots$. {\bf Step 8.} Next we proceed to show that the one-dimensional operators $P_{B_i, B_j}(x)= \langle B_i, x \rangle B_j$, where $B_i$, $B_j$ are basis elements, i.e., elements of the set $\{E_n, F_k, G_k: n \geq 0, k \leq -1\}$, also belong to $\mathcal I$. Since $z^m P_{E_n}E_n= E_{n+m}$ we see that $P_{E_n, E_{n+m}}=z^m P_{E_n}$ for every $n,m \geq 0$. Similarly we observe that $P_{E_n, E_{n-m}}=(z^*)^mP_{E_n}$ for $m \leq n$. Together, this proves that all operators $P_{E_n, E_k} \in \mathcal I$ for any $n,k \geq 0$. Next, we observe that \begin{gather*} z^m P_{F_n}= \begin{cases} r_1^m F_{n+m} & \text{if}\quad n+m <0, \\ r_1^{-n}E_{n+m} & \text{if}\quad n+m \geq 0. \end{cases} \end{gather*} Moreover, $(z^*)^m P_{F_n}=r_1^m F_{n-m}$ for all $n<0, m\geq 0$. Consequently, $P_{F_n,F_{n+m}}=r_1^{-m}z^m P_{F_n}$ and $P_{F_n,F_{n-m}}=r_1^{-m} (z^*)^m P_{F_n}$. Hence, $P_{F_n, F_k} \in \mathcal I$ for all $n,k \leq -1$. Moreover, $P_{F_n,E_k}=r_1^nz^mP_{F_n}$ and so $P_{F_n,E_k}\in\mathcal I$ for $k\geq 0$, $n\leq -1$. In fact, it can be easily verif\/ied that, $P_{B_j, B_i}^*=P_{B_i, B_j}$. This would mean that $P_{E_k,F_n}$ also belong to $\mathcal I$. Similar calculations show that $P_{G_n, G_{n+m}}= r_2^{-m}P_{G_{n+m}}z^m P_{G_n}$ for $n+m<0$ and $P_{G_n, G_{n-m}}= r_2^{-m}P_{G_{n-m}}(z^*)^m P_{G_n}$ for $n<0, m\geq 0$. Collectively, these imply that $P_{G_n, G_k} \in \mathcal I$ for all \mbox{$n,k \leq -1$}. Also $P_{G_n, E_{n+m}}= r_2^{m}P_{E_{n+m}}z^m P_{G_n}$ for $n+m \geq 0$. Hence $P_{G_n, E_k}$ and $P_{E_k,G_n}$ belong to $\in \mathcal I$ for all $k \geq 0$, $n \leq -1$. Finally, we notice that $P_{G_n, F_{-m}}= r_1^{-m}r_2^n P_{F_{-m}}(z^*)^m P_{E_0}z^n P_{G_n}$, which shows that the operators $P_{G_n, F_k} \in \mathcal I$ for all $n,k \leq -1$. Consider now f\/inite rank operators which are f\/inite linear combinations of the one-dimensio\-nal~$P_{B_i, B_j}$ for~$B_i$,~$B_j$ in the basis for~$H$. It is a~simple exercise in functional analysis to show that all compact operators are norm limits of such f\/inite rank operators. \end{proof} To describe the structure of $C^*(z)$ we need to understand the commutative quotient $C^*(z)/\mathcal{K}$. For the case of quantum pair of pants that structure and the idea of proof is very similar to the case of quantum annulus. Below we will show that the $C^*$-algebra $C^*(z)$ contains some inf\/inite-dimensional projections. Those are obtained from the spectrum of~$zz^*$. While tedious, the computation of the spectrum of~$zz^*$ is fairly straightforward and it amounts to studying a~(multi parameter) system of two step dif\/ference equations with constant coef\/f\/icients. The results of the computations are presented in the next three theorems. We start with the computation of the pure-point spectrum. \begin{Theorem} \label{discspecthm} The operator $zz^*$ has three eigenvalues: $1$~with eigenspace $\operatorname{span}\{E_n\}_{n\geq 0}$, $r_1^2$~with eigenspace $\operatorname{span}\{F_n\}_{n< 0}$, and the simple eigenvalue $\frac{r_1^2(a^2-r_2^2-r_1^2)}{a^2-r_1^2}$. \end{Theorem} \begin{proof} We study $(zz^*-\lambda)\varphi=0$. Using Lemma~\ref{z_z_star_coefficients} we get the following system of equations \begin{alignat}{3} & (1-\lambda)e_n=0 && \text{for}\quad n\ge1,&\nonumber \\ & \big(r_1^2+ r_2^2-\lambda\big)e_0+ar_2g_{-1}=0,&&&\nonumber \\ &\big(r_1^2-\lambda\big)f_n=0 &&\text{for}\quad n\le-1,& \label{kernelequ} \\ & ar_2g_{n+1}+\big(a^2+r_2^2-\lambda\big)g_n+ar_2g_{n-1}=0 \qquad && \text{for}\quad n\le-2,&\nonumber \\ & ar_2e_0+\big(a^2+r_2^2-\lambda\big)g_{-1}+ar_2g_{-2}=0.&&&\nonumber \end{alignat} The f\/irst and the third equations in this system yield the eigenvalues~1,~$r_1^2$ and the eigenspaces $\operatorname{span} \{E_n\}_{n\geq 0}$, $\operatorname{span}\{F_n\}_{n< 0}$ respectively. For the fourth equation, which is a~two step linear recurrence with constant coef\/f\/icients, the characteristic equation~is \begin{gather*} ar_2x^2+\big(r_2^2+a^2-\lambda\big)x+ar_2= 0. \end{gather*} The discriminant~$\Delta$ of this equation is \begin{gather*} \Delta=\big(r_2^2+a^2-\lambda\big)^2-4a^2r_2^2=\left((a-r_2)^2-\lambda\right)\left((a+r_2)^2-\lambda\right), \end{gather*} and the roots are \begin{gather*} x_{\pm}=\frac{\lambda-r_2^2-a^2 \pm \sqrt{\Delta}}{2ar_2}. \end{gather*} Thus, the formal solution to the homogeneous equation is \begin{gather} \label{ghomsol} g_n=c_+x_+^{n+1}+c_-x_-^{n+1}, \end{gather} where $c_+$ and $c_-$ are arbitrary constants. Notice that if we adopt the convention $g_0:=e_0$ in the last equation of~\eqref{kernelequ}, then this formula holds for $n\le 0$. There are three separate cases to consider: $0<\lambda<(r_2-a)^2$, $(r_2-a)^2\leq \lambda\leq (r_2+a)^2$ and $(r_2+a)^2<\lambda<1$. If $0<\lambda<(r_2-a)^2$ we notice the following facts about the solutions. Since $\lambda<(r_2-a)^2=r_2^2+a^2-2ar_2$ and both~$a$ and $r_2$ are positive, we have $\lambda-r_2^2-a^2<0$, and since $\Delta>0$, we have that $x_-<0$. Also notice that $x_+x_-=1$ from which it follows that $x_+<0$. Since $x_-<x_+$ we have $x_-<-1$ and $-1<x_+<0$. Thus, $|x_-|>1$ and $|x_+|<1$. Since we need $g_n\in\ell^2({\mathbb Z}_{<0})$, and since $|x_+|<1$, we must have $c_+=0$. In particular this implies that $g_0=x_-g_{-1}$. The second equation of the system~\eqref{kernelequ} then becomes \begin{gather*} \big((r_1^2+r_2^2-\lambda)x_-+ar_2\big)g_{-1} =0. \end{gather*} If $(r_1^2+ r_2^2-\lambda)x_-+ar_2=0$, then using the relation $x_+x_- =1$ we obtain \begin{gather*} x_+=\frac{\lambda-r_1^2-r_2^2}{ar_2}. \end{gather*} On the other hand we also have \begin{gather*} x_+=\frac{\lambda-a^2-r_2^2+\sqrt{\Delta}}{ar_2}. \end{gather*} Setting these equal to each other and solving for~$\Delta$ we get $\Delta=(\lambda-2r_1^2-r_2^2+a^2)^2$. Solving for~$\lambda$ yields \begin{gather*} \lambda=r_1^2\left(1-\frac{r_1^2}{a^2- r_1^2}\right)=\frac{r_1^2(a^2-r_2^2-r_1^2)}{a^2-r_1^2}. \end{gather*} First notice that due to the conditions on~$a$, $r_1$ and $r_2$ we have that $\lambda>0$ and $\lambda <r_1^2$ as $r_1^2/(a^2-r_1^2)>0$. Therefore this~$\lambda$ is in the interval $(0,(r_2-a)^2)$, and $c_-=g_{-1}$ is arbitrary. Consequently $\lambda =\frac{r_1^2(a^2-r_2^2-r_1^2)}{a^2-r_1^2}$ is a~simple eigenvalue with an eigenvector \begin{gather*} \varphi_\lambda=\sum\limits_{n=-\infty}^0x_-^{n+1}G_n. \end{gather*} In the case $(r_2+a)^2<\lambda<1$, we have $\lambda-r_2^2-a^2=\lambda-(r_2+a)^2+2r_2a>0$, $\Delta>0$ and hence $x_+>0$. Again from $x_+x_-=1$ we see that $x_+>1$ and $0<x_-<1$. Consequently, in equation~\eqref{ghomsol} we must have $c_-=0$, which then implies that $g_0=x_+g_{-1}$. The second equation of~\eqref{kernelequ} then becomes \begin{gather*} \big(\big(r_1^2+r_2^2-\lambda\big)x_++ar_2\big)g_{-1} =0. \end{gather*} Suppose there is~$\lambda$ such that $x_+(r_1^2+ r_2^2-\lambda)+ar_2=0$, which would then imply that \begin{gather*} x_+=\frac{ar_2}{\lambda-r_1^2-r_2^2} \end{gather*} and, since $x_+>1$, we must have \begin{gather*} \lambda < r_1^2+ar_2+r_2^2 < a^2+2ar_2+r_2^2=(r_2+a)^2. \end{gather*} This is a~contradiction. Consequently $g_{n}=0$ for every~$n$ and there are no eigenvectors in this case. When $(r_2-a)^2\leq\lambda\leq (r_2+a)^2$ the discriminant $\Delta\leq 0$ and the two solutions $x_+$ and $x_-$ are complex numbers conjugate to each other with absolute value equal to one. Since we need $g_n\in\ell^2({\mathbb Z}_{<0})$, equation~\eqref{ghomsol} implies that $g_{n}=0$ and there are again no eigenvectors in this case. \end{proof} The second, and the most technical step in the calculation of the spectrum of $zz^*$ is the calculation of the inverse of $zz^*-\lambda$. Norm estimates of the inverse provide insight about the resolvent set of $zz^*$. \begin{Lemma} \label{resolvent_z_z_star} The operator $(zz^*-\lambda)^{-1}$ is bounded for~$\lambda$ not an eigenvalue and $\lambda\in(0,(r_2-a)^2)\cup((r_2+a)^2,1)$. \end{Lemma} \begin{proof} We study $(zz^*-\lambda)\varphi=\tilde{\varphi}$ using coordinates \begin{gather*} \varphi=\sum\limits_{n=0}^\infty e_nE_n+\sum\limits_{n=-\infty}^{-1}(f_nF_n+g_nG_n) \qquad \text{and} \qquad \tilde{\varphi}=\sum\limits_{n=0}^\infty \tilde{e}_nE_n+\sum\limits_{n=-\infty}^{-1}\big(\tilde{f}_nF_n+\tilde{g}_nG_n\big). \end{gather*} Using Lemma~\ref{z_z_star_coefficients} we get the following system of equations \begin{alignat}{3} & (1-\lambda)e_n=\tilde{e}_n && \text{for}\quad n\ge1,&\nonumber \\ &\big(r_1^2+ r_2^2-\lambda\big)g_0+ar_2g_{-1}=\tilde{g}_0,&&&\nonumber \\ &\big(r_1^2-\lambda\big)f_n=\tilde{f}_n && \text{for}\quad n\le-1,&\nonumber \\ & ar_2g_{n+1}+\big(a^2+r_2^2-\lambda\big)g_n+ar_2g_{n-1}=\tilde{g}_n \qquad & &\text{for}\quad n\le-1.& \label{system_resolvent_z_z_star} \end{alignat} For the purpose of this proof we have introduced the notation $g_0:=e_0$ and $\tilde{g}_0:=\tilde{e}_0$. The f\/irst and the third equations in system~\eqref{system_resolvent_z_z_star} can be solved directly \begin{alignat*}{3 & e_n=\frac{1}{1-\lambda}\tilde{e}_n \qquad && \text{for}\quad n\ge1 \quad \text{and} \quad \lambda\neq 1, & \\ & f_n=\frac{1}{r_1^2-\lambda}\tilde{f}_n \qquad && \text{for}\quad n\le-1 \quad \text{and} \quad \lambda\neq r_1^2. & \end{alignat*} As for the fourth equation in system~\eqref{system_resolvent_z_z_star}, we start with the case $0<\lambda<(r_2-a)^2$. From the proof of Theorem~\ref{discspecthm} we have that $|x_-|>1$ and $|x_+|<1$, where $x_{\pm}=\frac{\lambda-r_2^2- a^2 \pm \sqrt{\Delta}}{2ar_2}$ are the solution of the characteristic equation $ar_2x^2+(r_2^2+a^2-\lambda)x+ar_2= 0$ with the discriminant $\Delta=((a-r_2)^2-\lambda)((a+r_2)^2-\lambda)$. We use the variation of parameters technique to solve this equation. For the homogeneous component of the solution, we use~\eqref{ghomsol} to obtain $g_n=c_+x_+^{n+1}+c_-x_-^{n+1}$ with $c_+$ and $c_-$ being constants to be determined. We look for the the solutions of the non-homogeneous equation in the form: $g_n=A_nx_+^{n+1}+B_nx_-^{n+1}$. Then the standard trick is to assume the f\/irst equation below to obtain the following system \begin{alignat*}{3} & (A_n -A_{n-1})x_+^{n+1}+(B_n-B_{n-1})x_-^{n+1}=0 \qquad &&\text{for}\quad n\leq 0,& \\ & (A_n -A_{n-1})x_+^n+(B_n-B_{n-1})x_-^n=\frac{-\tilde{g}_n}{ar_2} \qquad && \text{for}\quad n\leq -1. & \end{alignat*} In particular, \begin{gather} \label{acompini} (A_0 -A_{-1})x_++(B_0-B_{-1})x_-=0. \end{gather} The solution of the above system is \begin{gather} \label{acompgen} A_n-A_{n-1}=\frac{\tilde{g}_nx_-^{n+1}}{\sqrt{\Delta}}, \qquad B_n-B_{n-1}=\frac{-\tilde{g}_nx_+^{n+1}}{\sqrt{\Delta}}. \end{gather} In solving these dif\/ference equations we pay attention to square summability, making sure we only consider convergent expressions in powers of $x_\pm$. This leads to the following special solution of the non-homogeneous equation \begin{gather*} A_n=\frac{1}{\sqrt{\Delta}}\sum\limits_{j=-\infty}^n\tilde{g}_jx_-^{j+1} \qquad \text{and} \qquad B_n=\frac{1}{\sqrt{\Delta}}\sum\limits_{j=n+1}^{-1}\tilde{g}_jx_+^{j+1} \end{gather*} with $B_{-1}=0$, $n\leq -1$. Consequently the general solution is \begin{gather*} g_n=c_-x_-^{n+1}+\frac{1}{\sqrt{\Delta}}\left(\sum\limits_{j=n+1}^{-1}\tilde{g}_jx_+^{j-n}+\sum\limits_{j=-\infty}^n\tilde{g}_jx_-^{j-n}\right) \end{gather*} for $n<-1$, since we want $g_n\in\ell^2({\mathbb Z}_{<0})$, and since $|x_+|<1$ we must have $c_+=0$. For $n=-1$ we get \begin{gather*} g_{-1}=c_-+\frac{1}{\sqrt{\Delta}}\sum\limits_{j=-\infty}^{-1}\tilde{g}_jx_-^{j+1}, \end{gather*} and for $n=0$, using~\eqref{acompini}, we obtain \begin{gather*} g_{0}=c_- x_-+A_0x_++B_0x_-= c_- x_-+A_{-1}x_++B_{-1}x_-=c_- x_-+\frac{x_+}{\sqrt{\Delta}}\sum\limits_{j=-\infty}^{-1}\tilde{g}_jx_-^{j+1}. \end{gather*} Next we study the second equation of system~\eqref{system_resolvent_z_z_star}. Substituting the formulas for $g_{0}$ and $g_{-1}$ we compute \begin{gather*} c_-=\frac{\tilde{g}_0}{x_-(r_1^2+r_2^2-\lambda)+ar_2}-\left(\frac{{x_+(r_1^2+r_2^2-\lambda)+ar_2}}{x_-(r_1^2+r_2^2-\lambda)+ar_2}\right) \frac{1}{\sqrt{\Delta}}\sum\limits_{j=-\infty}^{-1}\tilde{g}_jx_-^{j+1}. \end{gather*} The above formulas give the unique solution of the equation $(zz^*-\lambda)\varphi=\tilde{\varphi}$, and hence def\/ine the inverse operator $(zz^*-\lambda)^{-1}$. We need to verify that this operator is bounded. So we f\/irst def\/ine the following operator \begin{gather*} Q\tilde{g}=\!\sum\limits_{n=-\infty}^0\!\! \left(\!\frac{\tilde{g}_0}{x_-(r_1^2+r_2^2-\lambda)+ar_2}- \!\left(\frac{{x_+(r_1^2+r_2^2-\lambda)+ar_2}}{x_-(r_1^2+r_2^2-\lambda)+ar_2}\right)\! \frac{1}{\sqrt{\Delta}}\! \sum\limits_{j=-\infty}^{-1}\!\! \tilde{g}_jx_-^{j+1}\!\right)\!x_-^{n+1}G_n. \end{gather*} Notice that, since $|x_-|>1$,~$Q$ is a~bounded operator taking $\ell^2({\mathbb Z}_{\leq 0})$ to itself. Here we have used the notation $G_0:= E_0$. Additionally we will need the following four operators, written in components: \begin{gather* \left(L_1g\right)_n =\frac{1}{\sqrt{\Delta}}\sum\limits_{j=-\infty}^{-1}x_-^{j-n}\chi\left(\frac{j}{n}\right)g_j: \ \ell^2({\mathbb Z}_{<0})\to \ell^2({\mathbb Z}_{<0}), \\ \left(L_2g\right)_n =\frac{1}{\sqrt{\Delta}}\sum\limits_{j=-\infty}^{-1}x_+^{j-n}\chi\left(\frac{n+1}{j}\right)g_j: \ \ell^2({\mathbb Z}_{<0})\to \ell^2({\mathbb Z}_{<0}), \\ \left(L_3g\right)_n =\frac{1}{\sqrt{\Delta}}\sum\limits_{j=-\infty}^{-1}x_+^{j-n}\chi\left(\frac{j}{n}\right)g_j: \ \ell^2({\mathbb Z}_{<0})\to \ell^2({\mathbb Z}_{<0}), \\ \left(L_4g\right)_n =\frac{1}{\sqrt{\Delta}}\sum\limits_{j=-\infty}^{-1}x_-^{j-n}\chi\left(\frac{n+1}{j}\right)g_j: \ \ell^2({\mathbb Z}_{<0})\to \ell^2({\mathbb Z}_{<0}). \end{gather*} Using these operators we write \begin{gather*} g_n=\left(Q\tilde{g}\right)_n+\left(L_1\tilde{g}\right)_n+\left(L_2\tilde{g}\right)_n \qquad\text{for}\quad n\leq -1 \end{gather*} and \begin{gather*} g_0 =\left(Q\tilde{g}\right)_0+x_+\left(L_1\tilde{g}\right)_{-1}. \end{gather*} Then we use Lemma~\ref{schuryounginq} (Schur--Young inequality) to estimate the norms of $L_1$ and $L_2$. We have \begin{gather*} \|L_1\|^2 \le \frac{1}{\Delta}\left(\underset{n\le-1}{\sup}|x_-|^{-n}\sum\limits_{j=-\infty}^n|x_-|^j\right) \left(\underset{j\le-1}{\sup}|x_-|^j\sum\limits_{n=j}^{-1}|x_-|^{-n}\right) \\ \phantom{\|L_1\|^2}{} =\frac{1}{\Delta}\left(\frac{1}{1-|x_-|^{-1}}\right)\left(\underset{j\le-1}{\sup}\frac{1-|x_-|^{j}}{1-|x_-|^{-1}}\right) =\frac{1}{\Delta}\frac{1}{(1-|x_-|^{-1})^2}. \end{gather*} The computation of norm of $L_2$ is similar. Therefore $(zz^*-\lambda)^{-1}$ is bounded for $0<\lambda<(r_2-a)^2$. The other case is when $(r_2+a)^2<\lambda<1$. From the proof of Theorem~\ref{discspecthm} we have that $x_+>1$ and $x_-<1$. Again using variation of parameters, we see that the particular solution of system~\eqref{acompgen} is given by: \begin{gather*} B_n=\frac{-1}{\sqrt{\Delta}}\sum\limits_{j=-\infty}^n\tilde{g}_jx_+^{j+1} \qquad \text{and} \qquad A_n=\frac{-1}{\sqrt{\Delta}}\sum\limits_{j=n+1}^{-1}\tilde{g}_jx_-^{j+1} \end{gather*} with $A_{-1}=0$, $n\leq -1$. Consequently the general solution is \begin{gather*} g_n=c_+x_+^{n+1} -\frac{1}{\sqrt{\Delta}}\left(\sum\limits_{j=n+1}^{-1}\tilde{g}_jx_-^{j-n}+\sum\limits_{j=-\infty}^n\tilde{g}_jx_+^{j-n}\right) \qquad \text{for} \quad n\leq-1. \end{gather*} Since we require $g_n\in\ell^2({\mathbb Z}_{<0})$, we must have $c_-=0$. For $n=-1$ we obtain \begin{gather*} g_{-1}=c_+-\frac{1}{\sqrt{\Delta}}\sum\limits_{j=-\infty}^{-1}\tilde{g}_jx_+^{j+1}, \end{gather*} and for $n=0$, using~\eqref{acompini}, we have \begin{gather*} g_{0}=c_+ x_++A_0x_++B_0x_-= c_+ x_++A_{-1}x_++B_{-1}x_-=c_+ x_+-\frac{x_-}{\sqrt{\Delta}}\sum\limits_{j=-\infty}^{-1}\tilde{g}_jx_+^{j+1}. \end{gather*} Substituting the formulas for $g_{0}$ and $g_{-1}$ into the second equation of system~\eqref{system_resolvent_z_z_star} we com\-pu\-te~$c_+$ \begin{gather*} c_+=\frac{\tilde{g}_0}{x_+(r_1^2+r_2^2-\lambda)+ar_2}+\left(\frac{{x_-(r_1^2+r_2^2-\lambda)+ar_2}}{x_+(r_1^2 +r_2^2-\lambda)+ar_2}\right) \frac{1}{\sqrt{\Delta}}\sum\limits_{j=-\infty}^{-1}\tilde{g}_jx_+^{j+1}. \end{gather*} The formulas above def\/ine the inverse operator $(zz^*-\lambda)^{-1}$. To verify that this operator is bounded we def\/ine the following operator \begin{gather*} R\tilde{g}=\!\sum\limits_{n=-\infty}^0\!\!\left(\!\frac{\tilde{g}_0}{x_+(r_1^2+r_2^2-\lambda)+ar_2}- \left(\frac{{x_-(r_1^2+r_2^2-\lambda)+ar_2}}{x_+(r_1^2+r_2^2-\lambda)+ar_2}\right)\! \frac{1}{\sqrt{\Delta}}\sum\limits_{j=-\infty}^{-1}\!\!\tilde{g}_jx_+^{j+1}\!\right)\!x_+^{n+1}G_n. \end{gather*} It is easy to see that~$R$ is a~bounded operator taking $\ell^2({\mathbb Z}_{\leq 0})$ to itself. Then we can write \begin{gather*} g_n=\left(R\tilde{g}\right)_n-\left(L_3\tilde{g}\right)_n-\left(L_4\tilde{g}\right)_n \qquad\text{for}\quad n\leq -1 \end{gather*} and \begin{gather*} g_0 =\left(R\tilde{g}\right)_0-x_-\left(L_3\tilde{g}\right)_{-1}. \end{gather*} We use Lemma~\ref{schuryounginq} to estimate the norms of $L_3$ and $L_4$, in a~manner similar to $L_1$. We omit the repetitive details. This shows that $(zz^*-\lambda)^{-1}$ is bounded for $(r_2+a)^2<\lambda<1$. \end{proof} In the theorem below, we will see that the interval $[(r_2-a)^2,(r_2+a)^2]$ is the continuous part of the spectrum of $zz^*$, completing its full description. \begin{Theorem} The spectrum of $zz^*$ is \begin{gather*} \sigma(zz^*)= \left\{\frac{r_1^2(a^2-r_2^2-r_1^2)}{a^2-r_1^2}\right\}\cup\big\{r_1^2\big\}\cup \big[(r_2-a)^2,(r_2+a)^2\big]\cup\{1\}. \end{gather*} \end{Theorem} \begin{proof} Since $zz^*$ is a~positive operator with norm 1, its spectrum must be a~closed subset of the interval $[0,1]$. In Theorem~\ref{discspecthm} we computed the pure point spectrum of $zz^*$ while Lemma~\ref{resolvent_z_z_star} identif\/ied intervals belonging to the resolvent set of $zz^*$. So it remains to analyze the interval $[(r_2-a)^2,(r_2+a)^2]$. We will show that if $\lambda\in((r_2-a)^2,(r_2+a)^2)$ then $\operatorname{Ran}(zz^*-\lambda)$, the range of $(zz^*-\lambda)$, is not all of~$H$. In the notation of system~\eqref{system_resolvent_z_z_star} consider $(zz^*-\lambda)\varphi=\tilde{\varphi}$ with $\tilde{g}_n=0$ for $n\leq -1$ but $\tilde{g}_0=\tilde{e}_0 \ne 0$. This leads to the equation \begin{gather*} ar_2g_{n+1}+\big(a^2+r_2^2-\lambda\big)g_n+ar_2g_{n-1}=0 \qquad\text{for}\quad n\leq -1, \end{gather*} with the general solution $g_n=c_+x_+^{n+1}+c_-x_-^{n+1}$. Since for $(r_2-a)^2\leq\lambda\leq (r_2+a)^2$, the two numbers $x_+$ and $x_-$ are complex conjugates each with magnitude one; which follows from the arguments in the last part of Theorem~\ref{discspecthm}; we must have $g_n =0$ for $n\leq 0$ for~$\varphi$ to be in~$H$. This however contradicts the second equation of~\eqref{system_resolvent_z_z_star} with $\tilde{g}_0=\tilde{e}_0 \ne 0$. Consequently there is no $\varphi\in H$ satisfying $(zz^*-\lambda)\varphi=\tilde{\varphi}$ for such $\tilde{\varphi}$ and $\operatorname{Ran}(zz^*-\lambda)$ is not~$H$. \end{proof} Let $P_E$, $P_F$, and $P_G$ be the orthogonal projections onto the inf\/inite-dimensional span of $\{E_n\}_{n\geq 0}$, $\{F_n\}_{n< 0}$ and $\{G_n\}_{n< 0}$ respectively. \begin{Proposition} The projections $P_E$, $P_F$, and $P_G$ belong to the noncommutative pair of pants, i.e., they are all in $C^*(z)$. \end{Proposition} \begin{proof} Since Lemma~\ref{resolvent_z_z_star} implies that $r_1^2$ is an isolated eigenvalue of $zz^*$, there exists a~continuous real valued function~$f$ so that $f(r_1^2)=1$ and~$f$ is zero on the rest of the spectrum of~$zz^*$. By functional calculus we have $f(zz^*)=P_F$ and so $P_F\in C^*(z)$. Similarly, since~$1$ is an isolated eigenvalue and we already know that $P_{E_0}\in C^*(z)$ we get $P_E\in C^*(z)$ as well. Since $P_G=I-P_E-P_F$, it then follows that $P_G\in C^*(z)$. \end{proof} We can decompose the Hilbert space~$H$ into $H \cong H_E \oplus H_F \oplus H_G$, where $H_E\cong\ell^2({\mathbb Z}_{\ge0})$, $H_F\cong\ell^2({\mathbb Z}_{<0})$, and $H_G\cong\ell^2({\mathbb Z}_{<0})$ are the Hilbert spaces with basis elements $E_n$, $F_n$, and $G_n$ respectively. Since $\ell^2({\mathbb Z}_{\ge0})$ is a~subspace of $\ell^2({\mathbb Z})$, which can be identif\/ied with $L^2(S^1)$ via the Fourier transform, we can view $H_E$ as a~subspace of $L^2(S^1)$. More precisely, if $B_n=\zeta^n$, $n\in{\mathbb Z}$ is the standard basis in $L^2(S^1)$, then $H_E$ is identif\/ied with the subspace $\operatorname{span}\{B_n\}_{n\geq 0}$ via \begin{gather*} E_n\mapsto B_n. \end{gather*} Def\/ine $P_{\ge0}: L^2(S^1)\to H_E$ to be the projection onto $H_E$. For a~$\varphi\in C(S^1)$ we def\/ine $T_E(\varphi):H_E \to H_E$ by $T_E(\varphi)=P_{\ge0}M(\varphi)$ where $M(\varphi):L^2(S^1)\to L^2(S^1)$ is the multiplication operator by~$\varphi$. In particular for $\varphi(\zeta)=\zeta$ we have{\samepage \begin{gather} \label{teone} T_E(\zeta)E_n=E_{n+1}, \end{gather} the unilateral shift.} Similarly we identify $H_F$ and $H_G$ with the subspace $\operatorname{span}\{B_n\}_{n< 0}$ in $L^2(S^1)$, and let $P_{<0}:L^2(S^1)\to H_F$ and $P_{<0}':L^2(S^1)\to H_G$ be the orthogonal projections onto $H_F$ and $H_G$ respectively. Then for a~$\varphi\in C(S^1)$, we def\/ine $T_F(\varphi):H_F \to H_F$ by $T_F(\varphi)=P_{<0}M(\varphi)$ and $T_G(\varphi):H_G\to H_G$ by $T_G(\varphi)=P_{<0}'M(\varphi)$ respectively. We have \begin{gather} \label{tfonf} T_F(\zeta)F_n=F_{n+1}, \qquad T_G(\zeta)G_n=G_{n+1} \qquad \text{for} \quad n<-1 \end{gather} and \begin{gather} \label{tgong} T_F(\zeta)F_{-1}=0, \qquad T_G(\zeta)G_{-1}=0. \end{gather} The operators $T_E(\varphi)$, $T_F(\varphi)$, and $T_G(\varphi)$ may be viewed as Toeplitz operators and they will be needed in proving the following result. \begin{Theorem} \label{quotient_algebra} The quotient $C^*(z)/\mathcal{K}$ is isomorphic to $C(S^1)\oplus C(S^1)\oplus C(S^1)$. \end{Theorem} \begin{proof} Using the above notation def\/ine \begin{gather*} T: \ C\big(S^1\big)\oplus C\big(S^1\big)\oplus C\big(S^1\big) \to C^*(z)/\mathcal{K} \end{gather*} by \begin{gather} \label{pants_toeplitz_def} T(\varphi_1,\varphi_2,\varphi_3)=T_E(\varphi_1)P_E+T_F(\varphi_2)P_F+T_G(\varphi_3)P_G+\mathcal{K} \end{gather} for continuous functions $\varphi_1$, $\varphi_2$, and $\varphi_3$ on the unit circle. To see that~$T$ is well def\/ined we need to show that $T(\varphi_1,\varphi_2,\varphi_3)$ is in $C^*(z)$. We showed that $P_E\in C^*(z)$, and notice that $T_E(\zeta)\in C^*(z)$ because $T_E(\zeta)=zP_E$. But Toeplitz operators $T_E(\varphi)$ can be uniformly approximated by polynomials in $T_E(\zeta)$ and its adjoint, and so $T_E(\varphi_1)P_E\in C^*(z)$. Similar arguments work for $T_F(\varphi_2)P_F$ and $T_G(\varphi_3)P_G$. We verify that~$T$ in~\eqref{pants_toeplitz_def} is a~isomorphism between the two algebras. First notice that equation~\eqref{pants_toeplitz_def} implies that~$T$ is continuous and linear. Next we show that the kernel of~$T$ is trivial. Consider the equation $T(\varphi_1,\varphi_2,\varphi_3)=0$ implying that $T_E(\varphi_1)P_E+T_F(\varphi_2)P_F+T_G(\varphi_3)P_G$ is compact. Since $P_E$, $P_F$ and $P_G$ are orthogonal, $T_E(\varphi_1)P_E$, $T_F(\varphi_2)P_F$ and $T_G(\varphi_3)P_G$ must be compact, and consequently $T_E(\varphi_1):H_E \to H_E$, $T_F(\varphi_2):H_F \to H_F$, and $T_G(\varphi_3):H_G \to H_G$ are compact. By the proof of Theorem~\ref{toeplitz_disk}, it follows that $\varphi_1=\varphi_2=\varphi_3=0$ and thus the kernel of~$T$ is trivial. Next we show that~$T$ is a~homomorphism of algebras. Consider the dif\/ference: \begin{gather} T(\varphi_1,\varphi_2,\varphi_3)T(\psi_1,\psi_2,\psi_3)-T(\varphi_1\psi_1,\varphi_2\psi_2,\varphi_3\psi_3) \nonumber \\ \qquad{} =\left(T_E(\varphi_1)T_E(\psi_1)-T_E(\varphi_1\psi_1)\right)P_E \nonumber \\ \qquad \phantom{=}{} +\left(T_F(\varphi_2)T_F(\psi_2)-T_F(\varphi_2\psi_2)\right)P_F+\left(T_G(\varphi_3)T_G(\psi_3)-T_G(\varphi_3\psi_3)\right)P_G. \label{homo_T} \end{gather} Since $T_E$, $T_F$, and $T_G$ are Toeplitz operators, the proof of Theorem~\ref{toeplitz_disk} implies that all three dif\/ferences on the right hand side of equation~\eqref{homo_T} are compact operators. Thus~$T$ is a~homomorphism between the two algebras. To show that the range of~$T$ is dense we consider the dif\/ference $T(\zeta,r_1\zeta, r_2\zeta+a)-z$. Using formulas~\eqref{teone},~\eqref{tfonf}, and~\eqref{tgong} we get $T_E(\zeta)E_n=E_{n+1}$, for $n\geq 0$, $T_F(r_1\zeta)F_n=r_1F_{n+1}$, for $n<-1$, and $T_G(r_2\zeta+a)G_n=r_2G_{n+1}+aG_n$, for $n<-1$. Observe that $T(\zeta,r_1\zeta, r_2\zeta+a)-z$ is not zero on $F_{-1}$ and $G_{-1}$ only and hence it is a~compact operator. Thus we have constructed functions $\varphi_1$, $\varphi_2$, and $\varphi_3$ such that $T(\varphi_1,\varphi_2,\varphi_3) = z$ in $C^*(z)/\mathcal{K}$. Since the $C^*$-algebra $C^*(z)/\mathcal{K}$ is generated by (the class of)~$z$, the range of~$T$ is dense and since the range of a~$C^*$-morphism must be closed,~$T$ is an isomorphism of algebras. This completes the proof. \end{proof} Note that from Theorem~\ref{quotient_algebra} we get a~short exact sequence \begin{gather*} 0\rightarrow \mathcal{K}\rightarrow C^*(z)\rightarrow C\big(S^1\big)\oplus C\big(S^1\big)\oplus C\big(S^1\big)\rightarrow 0. \end{gather*} We can compare this to the short exact sequence for the classical pair of pants \begin{gather*} 0\rightarrow C_0(PP_{(a, r_1,r_2)})\rightarrow C(PP_{(a, r_1,r_2)})\rightarrow C\big(S^1\big)\oplus C\big(S^1\big)\oplus C\big(S^1\big)\rightarrow 0, \end{gather*} where $C_0(PP_{(a, r_1,r_2)})$ are the continuous functions on the pair of pants that vanish on the boundary. \pdfbookmark[1]{References}{ref}
\section{Introduction} Spontaneous parametric down-conversion (SPDC) is a nonlinear optical process where a photon from a pump beam passing through a transparent crystal spontaneously splits into two daughter photons. The SPDC process has received continued interest because the daughter photons can be quantum mechanically entangled in the photon's degrees of freedom under appropriate conditions \cite{Shih}. The entanglement quality can be very high, enabling a wide range of experiments, from testing the foundations of quantum mechanics to applications in quantum communication, for example. Given that the SPDC process is an important resource for quantum information science, several groups have explored methods for optimizing different performance metrics of SPDC sources. The range of accessible experimental parameters is quite large, so many of these studies restrict their analysis to a subset of parameter space or resort to numerical simulations, which makes it difficult to easily generalize the results of these studies to different situations. In greater detail, the SPDC process most often involves the interaction of a pump beam (frequency $\omega_p$, propagation wavevector $\vec{k}_p$), and two generated beams, often called signal ($\omega_s$, $\vec{k}_p$) and idler ($\omega_i$, $\vec{k}_i$) beams, propagating through a transparent birefringent crystal characterized by a second-order nonlinear optical susceptibility. Conservation of energy requires that $\omega_p=\omega_s+ \omega_i$, and the SPDC process occurs most efficiently when momentum is conserved among the incident and generated beams - often referred to as phase matching - so that $\vec{k}_p=\vec{k}_s+\vec{k}_p$. In some situations, it is desirable to spatially separate the signal and idler beams and couple them into single-mode optical fibers for subsequent processing and transmission. In this paper, we consider a SPDC source consisting of a thin nonlinear optical crystal using the so-called Type-I interaction. Here, the all the beams are linearly polarized, the signal and idler beams are co-polarized, and the pump beam polarization is orthogonal to the generated beams. Phase matching is obtained by tuning the angle of the beams with respect to the crystal optic axes. In our study, we assume that the pump beam is in the fundamental Gaussian spatial mode, and the signal and idler beams are emitted collinearly or at a small angle with respect to the pump beam, shown schematically in Fig.~\ref{pumpgeo}. This configuration is a basic building block for realizing a bright source of polarization-entangled photons \cite{Kwiat95, Kwiat99}, for example. Subsequently, the generated beams are coupled into single-mode fibers, spectrally filtered, and directed to photon-counting detectors. \begin{figure} \begin{center} \includegraphics[scale=0.4]{pumpgeowithrefraction.pdf} \end{center} \caption{\textbf{SPDC interaction geometry} (a) A pump beam (blue) is focused at the center of a nonlinear optical crystal of length $L$. The fixed target modes (red) for the signal (idler) beams are defined by single-mode optical fibers, which are back-imaged to the center of the crystal. (b) Relation among the beam wavevectors and the phase mismatch. } \label{pumpgeo} \end{figure} We focus on two metrics of the SPDC source: the \textit{joint rate} (also known as the coincidence or biphoton rate) $R$, which is the rate of correlated photon events in the signal and idler beam paths, and the symmetric \textit{heralding efficiency} $\eta=R/\sqrt{R_s R_i}$ where $R_s$ ($R_i$) is the rate of single photon detections in the signal (idler) path. Optimizing the joint rate is crucial for quantum key distribution systems, for example, because the secure key rate is proportional to $R$ \cite{RevModPhysGisin}. The heralding efficiency takes on a value less than one when a photon is detected in the signal path, for example, but the correlated photon does not appear in the idler path. Enhancing $\eta$ is crucial for a number of applications that require a minimum value, including detection loophole-free tests of Bell's inequality ($\eta \geq 66\%$) \cite{loopholefree, bradprl, zeilenger}, one-sided device-independent QKD (DI-QKD) ($\eta \geq 66\%$) \cite{DIQKD}, and three-party quantum communication \cite{Pingpong1, Pingpong2} ($\eta \geq 60\%$). Also, high heralding efficiency decreases the quantum bit error rate and hence increases the error correction efficiency for point-to-point quantum key distribution systems. Both the joint rate and the heralding efficiency depend on most of the accessible parameters and there often exists a trade-off between the two: obtaining high $\eta$ often entails a reduction in $R$ \cite{Bennink}. In particular, we find that geometry (collinear or slightly noncollinear), frequency of the generated photons, beam spot sizes, and spectral filtering can all can have a important effect on $R$ and $\eta$. We motivate and give a physical explanation for why high $R$ and $\eta$ can be obtained in some configurations and not in others by investigating how the accessible parameters affect both the joint and and signal/idler spectra. We find that $\eta > 97\%$ can be obtained for the case of collinear SPDC when the frequencies of the signal and idler beams do not overlap so that they can be spatially separated with a dichroic mirror. Furthermore, we show $R$ and $\eta$ increase with overall decrease in mode waist sizes of pump, signal, and idler modes. For noncollinear SPDC, high $\eta$ can only be obtained with spectral filtering, which decrease $R$. We verify selected predictions in an experiment where we pump a thin BiB$_3$O$_6$ (BiBO) crystal pumped by an ultraviolet modelocked laser. The paper is organized as follows: In Sec.~II we derive both the joint spectrum and the singles spectrum of the SPDC photons and show how the dependence on the phase mismatch makes spectra different for different geometries and waists. In Sec.~III we discuss the heralding efficiency and joint count rate in both spectrally filtered and unfiltered cases and discuss its dependence on overall scaling of the mode waists as well as relative scaling of mode waists. In Sec.~IV we discuss several parameters that optimize heralding efficiency for all cases and briefly discuss how to optimize both heralding efficiency and joint counts for each geometry. In Sec.~V we present our experimental setup, discuss our results, and compare them to our theoretical findings and in Sec.~VI we conclude. \section{Formalism for Predicting the Joint and Singles Spectral Rates} The procedure for predicting the joint and singles spectral rates for SPDC into single-mode is well established and can be found across several studies. In our work, we closely follow and use similar notation for the joint spectral rate as described in Ling \textit{et al.} \cite{Ling}, who consider noncollinear SPDC, but do not predict the singles spectra or counts, which is needed to predict $\eta$.\footnote{We note that Eq.~(32) of Ling \textit{et al.} is in error, limiting the remainder of their work to Type-II SPDC or to non-degenerate Type-I SPDC. We do not use any results from Ling \textit{et al.} beyond Eq.~(32) so that our results are applicable for the Type-I interactions considered here.} For the singles spectral rates, we adapt the approach of Bennink \cite{Bennink}, who only considers collinear SPDC geometries. Finally, we adapt the formalism of Mitchell \cite{Mitchell} to include spectral filtering. Briefly, the procedure involves the quantum mechanical interaction Hamiltonian for the SPDC process, which is used to predict the transition rate for populating the initially empty signal and idler single-spatial modes (so-called `target' modes) with a biphoton state under the assumption of weak conversion so that perturbation theory applies \cite{MandelWolf}. The interaction Hamiltonian involves an integral of the product of the pump, signal and idler modes over the volume of the nonlinear optical crystal and accounts for the phase mismatch between the beams. We assume that the paraxial approximation holds for all beams, the pump beam is in a coherent state and that the transverse spatial profiles of the pump beam as well as modes collected by the optical fibers are given by a lowest-order Gaussian functions \begin{equation} U_{j}(\vec{r}) = e^{-(x_{j}^{2}+y_{j}^{2})/W_{j}^{2}}~(j=p,s,i), \end{equation} where $W_{j}$ is the 1/$e$ radius of the field mode. Furthermore, we assume that the transverse extent of the crystal is large enough that it does not cause diffraction of these modes, the crystal is thin enough so that the Guoy phases of the Gaussian beams are constant over the length of the crystal, birefringent walk-off can be ignored for the thin crystal, and that the pump beam is monochromatic so that there is a definite relation between the signal and idler frequencies via the energy conservation relation. The fiber collection modes, assumed to lie in the $x-y$ plane and described using the primed coordinate systems shown in Fig.~\ref{pumpgeo}, are back-propagated from the fiber, through the imaging lenses, and to the nonlinear optical crystal, where they experience refraction at the air-crystal interface. In general, the angle of refraction is frequency dependent due to the crystal dispersion, but we find the spectral rates are modified only slightly from this effect. Therefore, for simplicity, we take the angle of refraction to be constant at the value determined from Snell's law using the crystal refractive index at the central frequencies of the signal and idler beams, resulting in target modes in the crystal described by the unprimed coordinate systems shown in Fig.~\ref{pumpgeo}, which can be transformed to the pump-beam coordinate system in terms of the signal (idler) emission angle $\theta_{s(i)}$ inside the crystal. The expressions for the spectral rates involve a factor related to the geometrical overlap of the spatial mode envelopes (see the Eq. (21), Fig.~2, and the related discussion in Ling \textit{et al.} \cite{Ling}), reducing the rates when the beams do not overlap. This effect is has only a minor impact ($<2\%$) on our predictions for the small emission angles considered here and for the tightest focusing conditions used in the simulations and experiments described below. To support our goal of obtaining analytic predictions that can be interpreted readily, we ignore this effect. \subsection{Joint Spectral Rate} Following this procedure and under the various assumptions described above, we find that the joint spectral rate is given by \cite{Ling} \begin{equation} \frac{dR}{d\omega_{s}}=\frac{\eta_s \eta_i Pd_{\textrm{eff}}^2\alpha_{s}^2\alpha_{i}^2\alpha_{p}^2\omega_{s}\omega_{i}}{\pi \epsilon_{0}c^3 n_{s} n_{i} n_{p}}|\Phi(\Delta \vec{k})|^2, \label{jointspectralrate} \end{equation} which is the rate of biphoton production per unit frequency interval of the signal beam. Here, $\eta_{s(i)}$ is the overall efficiency of the signal (idler) paths, including any losses due Frensel reflections, absorption, and the detector quantum efficiency, $d_{\textrm{eff}}$ is the effective second-order nonlinear coefficient for the crystal, $P$ is the average pump power, $\epsilon_{0}$ is the permitivity of free space, $c$ is the speed of light in vacuum, and $\alpha_{j} = \sqrt{2/(\pi W_j^2)}$, are the Gaussian mode normalization constants. In Eq.~\ref{jointspectralrate}, the fact that the differential of $R$ with respect to frequency is in terms of the signal frequency $\omega_s$ reflects our assumption that the pump-beam frequency is monochromatic and hence $|d\omega_s|=|d\omega_i|$. The efficiency function in Eq.~\ref{jointspectralrate} is given by \begin{equation} \Phi(\Delta \vec{k}) = \frac{\pi L}{\sqrt{AC}}e^{-{\Delta k_{y}^{2}/(4C)}}\,\mathrm{sinc}\left(\frac{\Delta k_{z}L}{2}\right). \label{phijoint} \end{equation} Here, \begin{equation} A = \frac{1}{W_{p}^{2}}+ \frac{1}{W_{s}^{2}}+ \frac{1}{W_{i}^{2}}, \end{equation} and \begin{equation} C = \frac{1}{W_{p}^{2}}+ \frac{\cos^{2}\theta_{s}}{W_{s}^{2}}+ \frac{\cos^{2}\theta_{i}}{W_{i}^{2}}. \end{equation} The phase mismatch is defined through the relations \begin{eqnarray} \Delta \vec{k} & = & \vec{k}_{p}-\vec{k}_{s}-\vec{k}_{i} \\ & = & \Delta k_y \hat{y} + \Delta k_z \hat{z}, \end{eqnarray} with $\vec{k}_{j} = n_{j}(\omega_j,\vec{k}_j)\omega_{j}\hat{k}_{j}/c $, where $n_{j}$ is the frequency- and angle-dependent refractive index, $\hat{k}_{j}$ is the propagation unit vector for mode $j$, and $\hat{x},\hat{y}$ are the unit vectors for the pump-beam coordinate system \cite{Beouff}. We assume that the wavevectors lie in the $y-z$ plane and that $k_p$ is in a plane containing two of crystal axes for the case of the biaxial BiBO crystal considered here. See Fig.~\ref{pumpgeo}(b) for an illustration of the phase mismatch and its relation to the emission angles. In terms of the wavevector magnitudes \begin{eqnarray} \Delta k_z & = & k_p - k_s \cos \theta_s - k_i \cos \theta_i \label{zphasemismatch} \\ \Delta k_y & = & k_s \sin \theta_s - k_i \sin \theta_i. \label{yphasemismatch} \end{eqnarray} The total joint rate is found from the joint spectral rate and is given by \begin{eqnarray} R & = &\frac{\eta_s \eta_i Pd_{\textrm{eff}}^2\alpha_{s}^2\alpha_{i}^2\alpha_{p}^2\omega_{s}\omega_{i}}{\pi \epsilon_{0}c^3 n_{s} n_{i} n_{p}} \nonumber \\ & \times & \int_{-\infty}^\infty T_s(\omega_s)T_i(\omega_p-\omega_s) |\Phi(\Delta \vec{k})|^2\,d\omega_s, \label{totaljointrate} \end{eqnarray} where $T_{s(i)}$ is the intensity transmission function of spectral filters placed in the signal (idler) path. The phase mismatch (Eqs.~\ref{zphasemismatch} and \ref{yphasemismatch}) plays an important role in determining the rates, as can be seen in Eq.~\ref{phijoint}. We find that a physical explanation for the results presented below is more easily understood by considering a Taylor-series expansion of the wavevector magnitudes of the signal and idler beams with respect to frequency and truncating the series after second order. Following this approach, we have \begin{equation} k_{s(i)}(\omega) \simeq k_{s(i)0}+\frac{n_{g,s(i)}}{c}(\omega-\omega_{s(i)0})+\frac{1}{2}k''_{s(i)}(\omega-\omega_{s(i)0})^2, \end{equation} where $k_{s(i)0}=k|_{\omega=\omega_{s(i)0}}$ is the wavevector magnitude, $n_{g,s(i)}=c\, \partial k_{s(i)}/\partial \omega_{s(i)}|_{\omega=\omega_{s(i)0}}$ is the group index, and $k''_{s(i)}= (1/c)(\partial n_{g,s(i)}/\partial \omega_{s(i)}|_{\omega=\omega_{s(i)0}})$ is the group velocity dispersion parameter, all evaluated at the carrier frequency $\omega_{s(i)}$ of the signal (idler) beam. The carrier frequencies are chosen so that the interaction is perfectly phased matched ($\Delta \vec{k}=0$) for these frequencies. The target modes are chosen by adjusting the angle between the pump wave vector and the crystal axes, $\theta_p$, so that \begin{eqnarray} k_p - k_{s0} \cos \theta_s - k_{i0} \sin \theta_i & = &0 \\ k_{s0} \cos \theta_s - k_{i0} \cos \theta_i & = & 0. \end{eqnarray} For the case of frequency-degenerate down-conversion ($\omega_{s0}=\omega_{i0}=\omega_p/2$), the angles of emission are identical ($\theta_s=\theta_i \equiv \theta$), as well as the group indices ($n_{g,s}=n_{g,i} \equiv n_g$) and group velocity dispersion parameters ($k''_s=k''_i \equiv k''$) resulting in the relations \begin{eqnarray} \Delta k_z & = & - k'' \cos \theta \, (\omega_s-\omega_p/2)^2 \label{approxkz} \\ \Delta k_y & = & 2 n_g \sin \theta (\omega_s-\omega_p/2)/c. \label{approxky} \end{eqnarray} Hence, the longitudinal phase mismatch (the $z$-component) is quadratic in frequency\footnote{The error in Eq.~(32) in Ling \textit{et al.} \cite{Ling} stems from the fact that they assumed a linear relation between $\Delta k_z$ and frequency, which is not valid for frequency-degenerate Type-I SPDC as shown here.} and the transverse phase mismatch (the $y$-component) is linear in frequency, vanishing for collinear down-conversion ($\theta=0$). For nondegenerate down-conversion, both $\Delta k_z$ and $\Delta k_y$ are dominantly linear functions of frequency. These observations have important implications for the spectral bandwidth of the down-converted light as discussed in Sec.~III. \subsection{Singles Spectral Rate} We calculate the singles spectral rate in a similar manner as the joint spectral rate. Briefly, the singles spectral rate for the signal (idler) is given by the joint spectral rate, but for emission into \textit{any} emission direction for the idler (signal). This is accomplished formally by defining a generalized efficiency function (see Eq.~\ref{phijoint}) that determines the overlap of the pump and signal (idler) modes and the entire set of transverse modes of the idler (signal) \cite{Bennink,Uren2}. For this procedure, we use the complete set of orthonormal Hermite-Gauss modes with generalized beam normalization parameters for the signal beam given by \begin{equation} \alpha_{s}^{(n,m)}= \sqrt{\frac{2}{2^{n+m}n!m!\pi W_{s}^2}} \,; \end{equation} a similar definition for the idler beam parameter is obtain by substitution $s \rightarrow i$. We find that the singles spectral rate for the signal beam is given by \begin{equation} \frac{dR_s}{d\omega_{s}}=\sum_{n,m=0}^{\infty} \frac{\eta_s \eta_i P d_{\textrm{eff}}^2 (\alpha_s^{(n,m)})^2 \alpha_{i}^2\alpha_{p}^2\omega_{s}\omega_{i}}{\pi \epsilon_{0}c^3 n_{s} n_{i} n_{p}} |\Phi_s^{(n,m)}(\Delta \vec{k})|^2. \label{singlesspectralrate} \end{equation} The generalized mode efficiency functions $\Phi_s^{(n,m)}$ can be determined analytically, although the expressions are lengthy for large $(n,m)$. We find that all odd numbered modes in the $x$ direction ($n$ odd) are zero due to the fact that $\Delta k_{x} = 0$. It is instructive to analyze the lower-order mode contributions to develop physical intuition underlying the differences between the joint and singles spectral rates. In particular, we find that \begin{align} \Phi_s^{(0,1)}(\Delta \vec{k})&= \frac{ i\pi \sqrt{2}e^{-\Delta k_{y}^{2}/(4C)}}{W_{i}A^{3/2}\sqrt{C}} \nonumber \\ & \times \bigg\{\cos \theta_{i}\Delta k_{y}L\,\mathrm{sinc}\left[\frac{\Delta k_{z} L}{2}\right] \nonumber \\ &+(\cos \theta_{i}D+\sin \theta_{i}2C) \nonumber \\ & \times \frac{L}{\Delta k_{z}}\left(\cos\left[\frac{\Delta k_{z}L}{2}\right]-2\,\mathrm{sinc}\left[\frac{\Delta k_{z}L}{2}\right]\right)\bigg\}, \label{phi01} \end{align} and \begin{align} \Phi_s^{(2,0)}(\Delta \vec{k})&= \frac{2\pi}{\sqrt{AC}}\left(\frac{2}{AW_{i}^{2}}-1\right)\nonumber\\ &\quad \times e^{-\Delta k_{y}^{2}/(4C)}L{\rm sinc}\left[\frac{\Delta k_{z} L}{2}\right], \label{phi20} \end{align} where \begin{equation} D = \frac{\rm{sin}(2\theta_{s})}{W_{s}^{2}}-\frac{\rm{sin}(2\theta_{i})}{W_{i}^{2}}. \end{equation} Finally, the total singles rate for the signal is given by \begin{eqnarray} R_s & = & \sum_{n,m=0}^{\infty}\frac{\eta_s Pd_{\textrm{eff}}^2(\alpha_{s}^{(n,m)})^2\alpha_{i}^2\alpha_{p}^2\omega_{s}\omega_{i}}{\pi \epsilon_{0}c^3 n_{s} n_{i} n_{p}} \nonumber \\ & \times & \int_{-\infty}^\infty T_s(\omega_s)|\Phi_s^{(n,m)}(\Delta \vec{k})|^2\,d\omega_s, \label{totalsinglesrate} \end{eqnarray} Similar expressions can be obtained for the idler spectral and total rates by appropriately substituting $s \rightarrow i$. We find good convergence in our predictions when we truncate the sums in Eqs.~\ref{singlesspectralrate} and \ref{totalsinglesrate} above $n$ and $m$ above 10. \section{Spectral Rates for Different Configurations} In this section, we use the formalism developed above to predict the spectral rates for four different configurations. The base configurations for all results assume the use of a 600-$\mu$m-long BiBO crystal pumped with a modelocked 355-nm-wavelength laser with a $\sim$10-ps-long pulse duration and an average power of $P=1$ mW. The spectral width of the pump light is 0.013 nm (sech$^2$ pulse shape), which is much less than the spectral width of the down-converted light and hence the formalism developed above, which assumes a monochromatic pump beam, is applicable \cite{Castelletto}. We consider two different focusing conditions: ``loose'' focusing with $W_p=250$ $\mu$m and $W_s=W_i=100$ $\mu$m and ``tight'' focusing with $W_p=150$ $\mu$m and $W_s=W_i=50$ $\mu$m. We also consider the ``degenerate'' SPDC configuration where $\omega_{s0}=\omega_{i0}$ (corresponding to a central wavelength of 710 nm) and ``nondegenerate'' configuration where the signal (idler) frequency corresponds to a wavelength of 850 nm (609.6 nm). We take the signal and idler transmission/detection losses to be zero so that $\eta_s=\eta_i=1$. BiBO is a biaxial crystal that can be phase matched in our spectral range for the case when the pump beam experiences an angle-independent refractive index and the down-converted light experiences an angle-dependent refractive index. The nonlinear coefficient $d_\mathrm{eff}$ for BiBO is larger than the commonly used uniaxial crystal BBO when the pump beam makes an azimuthal angle of 90$^\circ$ with respect to the BiBO crystal axes and the polar angle is adjusted to obtain the desired SPDC geometry \cite{Hellwig, Ghotbi}. Figure~\ref{anglevswave} shows the angle of emission for the signal beam outside the crystal $\theta'_s$ that gives rise to perfect phase matching ($\Delta \vec{k}=0$) as a function of wavelength for BiBO and for different phase matching angles $\theta_p$. For nondegenerate SPDC, there are two solutions to the phase matching condition so that two spectral peaks are expected in both the signal and idler paths. These do not necessarily give rise to correlated photons and hence some care is needed to make sure that these spectral bands are separated and subsequently filtered. \begin{figure} \begin{center} \includegraphics[scale=.45]{anglevswavelengthnew.pdf} \caption{\textbf{Opening angle versus wavelength} Exterior opening angle as a function of wavelength for crystal tilt angles: $\theta_{p}=141.9^{\circ}$ (blue line), $\theta_{p}=142.2^{\circ}$ (maroon dashed line), and $\theta_{p}=143.0^{\circ}$ (gold line). The curves intersect this line in two places showing that both wavelengths will be detected in the fiber. The vertical dashed lines represent a conjugate pair of wavelengths. These show that the two angles in the same direction are similar to, but not conjugate wavelengths for noncollinear geometries.} \label{anglevswave} \end{center} \end{figure} \subsection{Degenerate Collinear SPDC} Here, we consider the case when the down-converted light is emitted in the same direction as the pump beams ($\theta_s=\theta_i=0$) and at the same frequencies. This configuration is not practical because it is not possible to separate the signal and idler beams, but it illustrates some important physics of the interaction. Figure~\ref{COLD}(a) shows the longitudinal phase mismatch as a function of frequency, where it is clearly seen that it is quadratic in frequency as predicted by Eq.~\ref{approxkz}. Furthermore, as seen in Eq.~\ref{approxky}, $\Delta k_y \sim 0$ (not shown). In this case, the exponential term in $\Phi(\Delta \vec{k})$ (Eq.~\ref{phijoint}) approaches unity and the efficiency is dominated by the sinc function; hence, the crystal length dictates the joint spectral rate shown in Fig.~\ref{COLD}(b) for the ``loose'' focusing condition. In particular, the width of the joint spectral rate can be determined approximately from the condition $\Delta k_z L \sim 2\pi$ and is independent of the focusing conditions. Given the quadratic nature of $\Delta k_z$, the joint spectral rate is quite broad and nearly constant about the central frequency. \begin{figure}[t] \begin{center} \includegraphics[scale=.38]{COLDdkzandspectra.pdf} \caption{\textbf{Collinear Degenerate} (a) The longitudinal phase mismatch as a function of frequency ($\omega_{s0} = 2.65\times10^{15}$ $s^{-1}$) for the collinear degenerate case. (b) The joint spectral rate (purple, dashed) and the singles spectral rate (blue, solid) are essentially equal. Here, $\theta_p=142^\circ$.} \label{COLD} \end{center} \end{figure} For the collinear case considered here, the higher-order spatial mode contributions ($n,m > 0$) to the singles spectral rates are very small. In particular, $D=0$ so that $\Phi_s^{(0,1)} \sim 0$ (see Eq.~\ref{phi01}). The frequency-dependent part of $\Phi_s^{(2,0)}$ arises from the sinc function and has an identical form as $\Phi$ so that the singles and joint spectral rates should have an identical frequency dependence. Furthermore, the scale factor in Eq.~\ref{phi20} can be made small using appropriate focusing. In particular \begin{equation} \left\lvert \frac{2}{A W_s^2}-1 \right\rvert=\frac{1}{1+2 W_p^2/W_s^2} \ll 1 \label{focusing} \end{equation} when $W_p \gg W_i$, where we have assumed that $W_s=W_i$, which is known to maximize $R$ \cite{Ling}. Hence, the higher-order-mode efficiency functions can be made small for the case when the pump is focused more loosely than the signal and idler modes. In this case, the singles spectral rate should be nearly identical to the joint spectral rate, as indeed is supported by the data shown in Fig.~\ref{COLD}(b). Thus, heralding efficiency could be very high (if the signal and idler beams could be spatially separated) because the spectral rates are essentially the same. The focusing condition for high heralding efficiency is opposed to the case for maximizing the total joint rate, which is optimized when the pump beam is focused tighter than the signal and idler modes \cite{Bennink,Ling,Dixon}. However, as we will see below, the count rate can be increased while maintaining or even improving the heralding efficiency by focusing all beams to a smaller waist but keeping the pump beam waist larger than the signal and idler waists. \subsection{Degenerate Noncollinear SPDC} To spatially separate the signal and idler beams for the case of degenerate SPDC, the crystal must be tilted by a small amount to achieve noncollinear phase matching, which results in the down converted light being emitted in a cone surrounding the pump beam with a small opening angle, typically 3$^\circ$ outside the crystal in a typical experiment. The correlated signal and idler beams are then located on opposite sides of the cone, as indicated in Fig.~\ref{pumpgeo}(a). For degenerate noncollinear down conversion, $\Delta k_z$ is essentially identical to that shown in Fig.~\ref{COLD}(a) (see Eq.~\ref{approxkz}), but $\Delta k_y$ becomes large and takes on a nearly linear frequency dependence as seen in Fig.~\ref{NONCOLD}(a) and predicted by Eq.~\ref{approxky}. For either of focusing conditions, the joint spectral rate is dominated by the transverse phase mismatch $\Delta k_y$ and the width of the joint spectral is given approximately by $\Delta k_y \sim 2 \sqrt{C}$. The parameter $C$ is inversely related to the beam waists and hence we expect that the joint spectral rate is now strongly dependent on the focusing conditions, being broader for tighter focusing, and largely independent of the crystal length. Figure ~\ref{NONCOLD}(b) and (c) show the joint spectral rate for the ``loose'' and ``tight'' focusing conditions, respectively, where tighter focus broadens the spectrum, as expected. Furthermore, because of the linear frequency dependence of $\Delta k_y$, the spectral shape is nearly Gaussian due to the exponential term in Eq.~\ref{phijoint}. \begin{figure} \begin{center} \includegraphics[scale=.38]{NONCOLDdkyandspectra.pdf} \caption{\textbf{Noncollinear Degenerate} (a) The transverse phase mismatch as a function of frequency for $\omega_{s0}=2.65\times10^{15}$ $s^{-1}$ for the noncollinear degenerate case. Joint spectral rate (purple, dashed) and singles spectral rate (blue, solid) for (b) ``loose'' and (c) ``tight'' focusing conditions. The green vertical lines at frequency offsets of $\pm 3.73 \times 10^{13} s^{-1}$ correspond to a full bandwidth of $\sim$20 nm for a central wavelength of 710 nm. Here, $\theta_p=141.9^\circ$, $\theta_s=\theta_i=1.64^\circ$, corresponding to $\theta'_s=\theta'_i=3.04^\circ$.} \label{NONCOLD} \end{center} \end{figure} The singles spectral rates are substantially different in comparison to the joint spectral rate for the noncollinear case, as seen in Figs.~\ref{NONCOLD}(b) and (c). This arises from the fact that a photon detected in the single mode of the signal path has its corresponding photon emitted into a higher-order mode in the idler path. In particular, $\Phi_s^{(0,1)}$ does not vanish and the term shown in the second line of Eq.~\ref{phi01} is linear in $\Delta k_y$, effectively broadening the spectrum. Indeed, in summing over all modes, we see in Fig.~\ref{NONCOLD}(b) that the singles spectral rate is much broader than the joint spectral rate and, by comparing panels (b) and (c), there is only a weak dependence in the singles spectral width for different focusing conditions. Inspection of panels (b) and (c) also reveals that tighter focusing gives rise to a substantially higher biphoton generation rate with the tighter focusing condition giving nearly a factor of 3 increase in the rate in comparison to the loose focusing condition. Based on this data, we anticipate that the heralding efficiency will be quite low for this noncollinear geometry, which will be explored quantitatively in Sec. IV. It is also clear why spectral filter can help improve the heralding efficiency. For a narrow-band spectral filter centered on the degenerate frequency, the integral of the joint spectral rate and the singles spectral rates (needed to determine $R$ and $R_{s(i)}$) over the filter bandwidth (vertical lines) can be made similar to each other. Furthermore, tighter focusing broadens the joint spectral rate while leaving the singles spectral rate approximately the same, so we expect higher heralding efficiency for this focusing condition. \subsection{Collinear Nondegenerate SPDC} We now consider the case of tilting the crystal to obtain collinear down conversion with different signal and idler frequencies. As in the collinear degenerate case, $\Delta k_y \simeq 0$, and the longitudinal phase mismatch is a nearly linear function of frequency, as seen in Fig.~\ref{COLND}(a), and hence the joint spectral rate has a width that is substantially narrower as seen in Fig.~\ref{COLND}(b). As in the degenerate collinear case, the shape of the spectrum is dictated by the sinc function in Eq.~\ref{phijoint} and hence depends mostly on the crystal length and is independent of the focusing conditions. Also, for the case when the pump waist is much larger than the target-mode waists, the singles spectral rate is essentially identical to the joint spectral rate (see Fig.~\ref{COLND}) for the same reasons as discussed above for the degenerate-frequency case. With this observation, we expect high heralding efficiency without the use of any spectral filtering. While not shown here, the magnitude of the spectral rates all increase for tighter focusing. \begin{figure} \begin{center} \includegraphics[scale=.38]{COLNDdkzandspectra.pdf} \caption{\textbf{Collinear Nondegenerate} (a) Longitudinal phase mismatch as a function of frequency. (b) The joint spectral rate (purple, dashed) and the singles spectral rates (green, solid) are essentially identical; there differences cannot be discerned. Here, $\omega_{s0}=3.09\times10^{15}$ $s^{-1}$ (corresponding to a wavelength of 850 nm), $\omega_{i0}=2.22\times10^{15}$ $s^{-1}$ (corresponding to a wavelength of 609.6 nm), and $\theta_p=143.22^\circ$.} \label{COLND} \end{center} \end{figure} \subsection{Noncollinear Nondegenerate SPDC} For the noncollinear situation, both the longitudinal and transverse phase mismatch play a role in determining the spectral rates. The longitudinal mismatch is essentially identical to the collinear case shown in Fig.~\ref{COLND}(a) for the small emission angles considered here, and the transverse phase mismatch is large and an approximately linear function of frequency as seen in Fig.~\ref{NONCOLND}. In this situation, the longitudinal phase mismatch is somewhat dominant so that the width of the joint spectral rate is only weakly dependent on the focusing conditions, as seen in Fig.~\ref{NONCOLND}(b) and (c), with a slight broadening for tighter focusing. Interestingly, the singles spectral rate is only slightly broader than the joint spectral rate so that the heralding efficiency should be higher in this case in comparison to the noncollinear degenerate case. Spectral filtering will also increase the heralding efficiency, although a much narrower filter is needed to accomplish this task given that the overall spectrum is narrower than for the degenerate case. As before, the overall rate is increased using the tighter focusing conditions. \begin{figure} \begin{center} \includegraphics[scale=.38]{NONCOLNDdkyandspectra.pdf} \caption{\textbf{Noncollinear Nondegenerate} (a) Transverse phase mismatch as a function of frequency for the noncollinear nondegenerate case. The joint spectral rate (blue, dashed) and the singles spectral rate for the signal (gold, solid) and idler (purple, dotted) for the (b) ``loose'' and (c) ``tight'' focusing conditions. The carrier frequencies are the same as those given in the caption to Fig.~\ref{COLND}, $\theta_p=142.44^\circ$, $\theta_s=3.05^\circ$, $\theta'_s=5.62^\circ$, $\theta_i=2.17^\circ$, and $\theta'_i=4.02^\circ$.} \label{NONCOLND} \end{center} \end{figure} \section{Heralding Efficiency and Joint Count Rates} As mentioned above, high heralding efficiency is obtained when the joint spectral rate is similar to the singles spectral rates, which can be forced to be more similar using filters centered on the signal and idler carrier frequencies. The ideal filter has a flat top profile of full width $\Delta \omega_f$ and unit transmission in the passband, which we assume in the plots in this section. We do not consider the collinear degenerate case further because of the difficulty in spatially separating the signal and idler beams. Figure~\ref{HE}(a) and (c) show the heralding efficiency and total joint rate for noncollinear degenerate SPDC as a function of filter bandwidth. The vertical line indicates a filter with a $\sim$20 nm bandwidth. It is seen that $\eta$ is quite low (below 60\%) for no spectral filtering (large $\Delta \omega_f$), but that the ``tight'' focusing condition has a substantially higher efficiency and rate for all filter bandwidths. Heralding efficiencies $>90\%$ can be obtained for the tighter focusing and a narrow filter bandwidth, but at the cost of the total joint rate. \begin{figure} \begin{center} \includegraphics[scale=.38]{newheraldingefficiency.pdf} \caption{\textbf{Heralding efficiency and joint counts versus filter bandwidth.} Purple (dashed) lines have waist parameters in the crystal of $W_{p} = 150 \mu m$, $W_{s} = W_{i} = 50 \mu \textrm{m}$ and blue (solid) lines have waist parameters of $W_{p} = 300 \hspace{2pt}\mu \textrm{m}$, $W_{s} = W_{i} = 100 \hspace{2pt}\mu \textrm{m}$. Vertical green lines at $7.46 \times 10^{13} s^{-1}$ show approximately where a 20 nm filter at 710 nm would be. (a) and (b) show the heralding efficiency as a function of filter frequency for the two sets of waist parameters for (a) noncollinear degenerate and (b) noncollinear nondegenerate. (c) and (d) show the joint count rates for (c) noncollinear degenerate and (d) noncollinear nondegenerate.} \label{HE} \end{center} \end{figure} For collinear nondegenerate SPDC, the heralding efficiency exceeds 99.5\% for both focusing conditions and without any spectral filtering. This is due to the similarity of the joint and singles spectral rates (Fig.~\ref{COLND}). For ``loose'' (``tight'') focusing, $R=19.5$ kHz/s/mW ($R=56.8$ kHz/s/mW). Hence, this configuration with tight focusing is very promising for obtaining high efficiency and rates if high transmission filters can be identified to block the copropagating pump light and for spatially separating the signal and idler beams. Noncollinear nondegenerate SPDC relaxes the constraints on the filters, but come at the price of a reduced $\eta$. As for the degenerate case, $\eta$ is higher for tighter focusing (see Fig.~\ref{HE}(b), but the overall efficiency is much higher than the degenerate case. Because the joint spectral rate is much narrower, a narrower band filter is needed to obtain the highest efficiencies, but $\eta>90\%$ can be obtained with a 20-nm-bandwidth filter with almost no loss in total rate (see Fig.~\ref{HE}(d)). Unfortunately, $R$ is substantially lower for the noncollinear nondegenerate rate because of the narrow spectrum in comparison to the noncollinear degenerate case with aggressive filtering, so the two geometries are nearly equivalent with respect to the efficiency and rate metrics. \section{Comparison to Experimental Findings} We compare our theoretical predictions to our experimental findings for the noncollinear degenerate case as well as the noncollinear nondegenerate case for a limited set of parameters. The pump beam is generated by a high-power modelocked laser (Coherent Palidan, 4 W maximum average power, 120 MHz pulse repetition rate, 10-ps-long pulse duration) focused into a 600-$\mu$m-long BiBO crystal (Newlight Photonics) anti-reflection coated at both the pump wavelength and the degenerate down conversion wavelength (710 nm). The down converted light is collected using a well-corrected achromatic lens (Schaefter and Kirchhoff GmbH, 60FC-T-0-M20l-02) and coupled into a single-mode fiber (Thorlabs custom fiber), that does not have anti-reflection coatings on the fiber ends for nondegenerate frequencies or anti-reflection coated fiber for degenerate frequencies (Oz Optics, custom fiber). To coarsely align the signal and idler paths, we back-propagate laser light through the fibers and lenses toward the BiBO crystal at wavelengths close to the emitted SPDC light, adjusting the spot size to the desired values and placing the waist in the crystal, a process that is enabled using a beam profiler (Thorlabs, BP209-VIS). Single photons are detected using silicon avalanche photodiodes operating in Geiger mode (Perkin-Elmer/Excelitas, SPCM - AQRH) with a peak quantum efficiency at 710 nm of $\sim 63\%$. The electrical pulses generated by the detectors are sent to a custom-programmed field-programmable gate array for coincidence counting ($\sim$9 ns coincidence window) and counting of single photon events. To measure the singles spectral rate, we couple the light from the fibers into a triple monochrometer (Newport, Cornerstone 260 1/4 m) and a photomultiplier tube (Hammamatsu, H6780) followed by transimpedance amplification and lock-in detection. The spectral response of the spectrometer system is calibrated using a high-pressure tungsten halogen lamp (Ocean Optics, HL2000). \textit{Noncollinear Degenerate} -- In light of the fact that the heralding efficiency is strongly dependent on the spectral rates, we measure the singles spectral rate for the signal beam, shown in Fig.~\ref{data}(a). It is seen that the emission is very broad band and is in excellent agreement with our predictions. Based on the discussion surrounding Fig.~\ref{NONCOLD}, spectral filter is required to obtain high heralding efficiency. We use a 23-nm-bandwidth filter (Semrock TBP 704/13) that is close to an ideal top hat with an efficiency of $~99\%$, which we angle tune to center the passband on the degenerate wavelength. Using the photon-counting setup, we measure $\eta=43\pm0.5\%$ with accidental coincidences contributing 0.5 $\%$ to this value (this value is not subtracted in the quoted efficiency) with $W_{p} = 250\pm5$ $\mu$m, $W_{s}= W_{i} = 100\pm5$ $\mu$m. To compare to our theoretical predictions, we need to correct for the transmission/detection losses, which we estimate as $\eta_s=\eta_i=75.2\pm1.2\%$), with contributions of $62.5\pm0.5\%$ for the detection efficiency, $84\pm2\%$ for the non-ideal behavior of the spectral filters and minimal loss on the AR-coated fibers of $<1\%$. Using these values, we obtain a corrected heralding efficiency $\eta_{correct}=81.7\pm2.6\%$ , which should be compared to our theoretical prediction of $82.1 \pm 1.3 \%$. The agreement is within our assigned errors and is excellent. To measure the total joint count rate, we adjust the average pump power to $\sim$100 mW so that the single count rates are well below the saturation rate of the detector and much larger than the dark count rate. We then scale the results to determine the rate for $P=1$ mW so we can directly compare to our predictions. We find that $R=982\pm20$ Hz. To find a corrected rate, we divide this result by the product $\eta_s \eta_i$ to arrive at $R_{correct}=2.39\pm0.05$ kHz, while we predict $R=2.16\pm0.13$ kHZ. The agreement is within approximately a standard deviation of our measurements, which is good. \begin{figure} \begin{center} \includegraphics[scale=.4]{datacollinearandnoncollinearwp250.pdf} \caption{\textbf{Predicted and measured spectra} Signal singles spectral rate for (a) noncollinear degenerate SPDC with $\theta'_s=3.04^{\circ}$ and (b) collinear nondegenerate SPDC. Here, the experimental data is shown by the blue solid lines and the predictions are shown by the red dashed lines, where we use a least-square fit to choose the vertical scaling of the theoretical predictions. The green bars show representative errors.} \label{data} \end{center} \end{figure} \textit{Noncollinear Nondegenerate} -- For the case of noncollinear nondegenerate SPDC with the signal (idler) central wavelength of 850 nm (609.6 nm) with the same waists as above we measure $\eta=29.4\pm0.6\%$, with accidental coincidences contributing 0.4$\%$. We do not use narrowband spectral filters for this measurement, but use a low-pass filter (Semrock FF01-650/SP-25, $>99.0\%$ transmission) in the idler arm to block out the 869.3-nm emission that arises from the second phase matching condition (see Fig.~\ref{anglevswave} and the associated discussion). We use a long-pass filter (Semrock BLP01-785R-25, $>99.0\%$ transmission) in the signal to block out the 600 nm arising for the same reason. We estimate that $\eta_s=34.9\pm1.4\%$ with contributions of $38\pm1.5\%$ from the detector and $8\pm0.2\%$ from the Fresnel reflections from the fibers, and $\eta_i=53.3\pm1.4\%$ with contributions of $58\pm1.5\%$ from the detector and $8\pm0.2\%$ from the Fresnel reflections from the fibers. Using these values, we find $\eta_{correct}=68.1\pm4.3\%$, which agrees very well with our predicted value of $71.6\pm1.2\%$ We measure $R=410\pm12$ Hz, which is corrected using the efficiencies given above to find $R_{correct}=1.89\pm0.19$ kHz, which compares favorably with the predicted joint count rate of $1.34\pm0.10$ kHz. \textit{Collinear Nondegenerate} -- Figure~\ref{data}(b) shows the measured and predicted singles spectral rate along the signal beam path, where it is seen that the spectral width is much narrower than that observed for degenerate SPDC. Also, both spectral peaks arise from the two solutions to the phase matching condition as discussed in relation to Fig.~\ref{anglevswave}. In the collinear case, the frequencies are conjugates of each other. The agreement with the predicted spectrum in both relative heights and spectral widths for both spectral features is excellent. To measure the heralding efficiency, we use use a single lens to collect both the signal and idler beams, couple into single mode fiber, and then back to free space where we use a dichroic mirror to separate the signal and idler beams, which are then coupled into multi-mode uncoated fibers and sent to the photon counting detectors. Because the achromat is not perfectly correct, the back-propagated waist at the crystal are different: $W_s=120\pm10$ $\mu$m and $W_i=95\pm10$ $\mu$m. The pump light is blocked using a high-pass filter (Semrock, Di02-R405-25x36 ) that has a transmission at the signal and idler wavelengths exceeding $95\%$ and a pump suppression greater than a factor of 10$^{6}$. No narrow-band spectral filters are placed in the signal or idler beam paths. We find that $\eta=33.5\pm0.5\%$, which is corrected to yield $\eta_{correct}=86\pm5\%$. Here, we estimate $\eta_s=29.2\pm1.8\%$ with contributions contributions of $38\pm1.5\%$ and a combined coupling loss of $77.5\% \pm1\%$ from coupling the beam into freespace and back to fiber and $98\%\pm1\%$ from the filter, and $\eta_i=53.3\pm2.3\%$ with contributions $58\pm1.5\%$ and a combined coupling loss of $92.8\% \pm0.7\%$ and $98\%\pm1\%$ from the filter. This result is considerably lower than the predicted value of $94.8\pm0.5\%$, where we have accounted for the different signal and idler waists in the theory. One possible reason for the lower measured heralding efficiency is that there is a shift in the location of the waist locations of the signal and idler beams of $\sim 250$ $\mu$m. This non-ideality can be avoided by first spatially separating the signal and idler beams and coupling them into independent single-mode fibers. We measure $R=392\pm12$ Hz for, which is corrected to yield $R_{correct}=2.45\pm0.19$ kHz, while the predicted joint count rate is $1.82\pm0.08$ kHz, where the agreement is reasonable. \textit{Tighter focusing} -- We also decreased the beam waists for the case collinear nondegenerate configuration, using $W_p = 150\pm5$ $\mu$m, $W_s = 67\pm5$ $\mu$m, and $W_{i} = 47\pm5$ $\mu$m. We find $\eta=32.2\pm0.6\%$, which is corrected to $\eta_{correct}=82.6\pm4\%$, while the theoretical prediction is slightly lowered for these beam parameters and is equal to $89\pm2\%$, which is slightly higher than the observations, although still nearly within error. Finally, we measure $R=890\pm45$ Hz, which is corrected to $R_{correct}=5.53\pm0.62$ kHz, which compares favorably with the predicted value of $4.81\pm0.55$ kHz. \section{Conclusions and Discussion} In conclusion, we develop a theoretical formalism to predict the joint spectral rate and the singles spectral rates for Type-I SPDC process using a thin nonlinear optical crystal. From these predictions, the physical factors influencing the heralding efficiency and total joint rate are identified, allowing us to design a system that has an intrinsically high efficiency and rate. We also find that the heralding efficiency depends on the focusing tightness for the noncollinear down-conversion configurations, especially for the degenerate case, and that the collinear configurations are less sensitive to this parameter. Finally, we obtain good agreement between theoretical predictions and experimental observations. We compare a few of our results to previous findings that optimize the SPDC heralding efficiency and count rate when the light is coupled into single mode optical fibers. Migdall \textit{et al.} (see Fig.~4, Ref. \cite{Migdall}) show the heralding efficiency increases as the signal waist, relative to the pump waist, decreases in the limit of a thin crystal, and that the joint count rate decreases with decreasing pump waist for both the collinear and noncollinear regimes. This is in agreement with what our theory predicts for pump waist scaling. This tradeoff between heralding efficiency is also in agreement with Bennink \cite{Bennink}, who finds that, as the pump waist is focused tighter, the heralding efficiency decreases, and the joint count rate increases. We agree generally with this result, although Bennink considers the thick-crystal regime. Recently, good agreement between experiments and Bennink's predictions has been reported by Dixon \textit{et al.} \cite{Dixon}. Baek and Kim \cite{Baek} show that both the joint and singles spectrum for frequency degenerate, Type-I collinear have a broad bandwidth, while for the noncollinear configuration, the singles spectrum remains broad, but the joint spectrum is much narrower, which is consistent with our findings. Finally, Carrasco \textit{et al.} \cite{Carrasco} show that the singles spectrum can be broadened in both the collinear and noncollinear configurations by decreasing the pump waist, and that the joint spectrum for noncollinear can also be broadened by tighter focusing of the pump beam. Our theory agrees with the results presented in Fig. 1 in \cite{Carrasco}. If we were to replace our detectors with high-quantum efficiency devices, such as recently develop WSi superconducting nanowire detectors with close to 100\% detection efficiency \cite{SaeWoo} and replace anti-reflection-coated fibers, heralding efficiencies over 80\% are possible for noncollinear degenerate SPDC, and potentially over 90\% for the collinear nondegenerate configuration if we were to solve the issue of the chromatic focusing of the signal and idler beams. Our work paves the way for optimized SPDC sources that will find applications in fundamental quantum information science as well as in practical quantum key distribution systems. \section*{Acknowledgment} The authors gratefully acknowledge financial support from the DARPA DSO InPho project and for extensive discussion on obtaining high heralding efficiency in SPDC with Paul Kwiat and Bradley Christensen. \ifCLASSOPTIONcaptionsoff \newpage \fi \bibliographystyle{IEEEtran}
\section{\label{}} \section{Introduction} The shell structure of nuclei established by Goeppert-Mayer \cite{Mayer49} and Haxel \textit{et al.} \cite{Haxel1949} more than 60 years ago is the corner stone of nuclear structure described by the shell model. However, a few decades later, with systematic studies of nuclei with large $N / Z$ ratio, known as "exotic nuclei", it was observed that the original shell gaps are not preserved and "new" shell closures appear \cite{Krucken2011,Sorlin2008,Wienholtz2013}. This fact continues to attract the attention of many experimentalists and theorists who try to understand the origin of these changes. Nowadays, despite the experimental challenges, a large variety of exotic nuclei can be produced and studied with highest precision in facilities around the world \cite{Blaum2013,VanDuppen2011}. These experimental data are used by theorists for fine tuning of the effective interactions in order to improve their descriptive as well as predictive power \cite{Heyde2013}. In the past decade, the region below Ca ($Z < 20$) with $ 20 \leq N \leq 28$ was investigated intensively, in particular the evolution of the $\pi sd$ orbitals as a function of neutron number (for review see e.g. Refs.\,\cite{Sorline2013,Gade2006}). The energy spacing between the $1/2^{+}$ and $3/2^{+}$ levels as a function of the $\nu f_{7/2}$ occupancy and the evolution of the $N = 20$ and $N = 28$ shell gaps with decreasing $Z$ for odd-A K ($Z = 19$), Cl ($Z = 17$) and P ($Z = 15$) was presented by Gade \textit{et al.} \cite{Gade2006}, with experimental results compared to shell-model calculations up to $N = 28$. The inversion of the $1/2^{+}$ and $3/2^{+}$ states in the Cl chain is observed for the half-filled $\nu 1f_{7/2}$ orbital. The same effect appears for potassium isotopes, but only when the same orbital is completely filled, at $N = 28$. In addition, the evolution of the effective single-particle energies (ESPE) for potassium isotopes (single-hole states in Ca isotopes) based on shell-model calculations is discussed by Smirnova \textit{et al.} in Ref.~\cite{Smirnova2012}, where a degeneracy of the $\pi 2s_{1/2}$ and $\pi 1d_{3/2}$ levels is predicted to occur at $N = 28$ and returns to a "normal" ordering ($\pi 2s_{1/2}$ below $\pi 1d_{3/2}$) approaching $N = 40$ (Fig\,1(c) in Ref.\,\cite{Smirnova2012}). The reordering of the orbitals is driven by the monopole part of the proton-neutron interaction, which can be decomposed into three components: the central, vector and tensor. Initially Otsuka {\it{et al.}} \cite{Otsuka2005} suggested that the evolution of the ESPEs is mainly due to the tensor component. However, in more recent publications \cite{Smirnova2010,Smirnova2012,Otsuka2010} several authors have shown that both the tensor term as well as the central term have to be considered. Regarding the shell model, potassium isotopes are excellent probes for this study, with only one proton less than the magic number $Z = 20$. Nevertheless, little and especially conflicting information is available so far for the neutron-rich potassium isotopes. Level schemes based on the tentatively assigned spins of the ground state were provided for $^{48}$K \cite{Krolas2011} and $^{49}$K \cite{Broda2010}. In addition, an extensive discussion was presented by Gaudefroy \cite{Gaudefroy2010} on the energy levels and configurations of $N = 27, 28$ and 29 isotones in the shell-model framework and compared to the experimental observation, where available. However, the predicted spin of $2^{-}$ for $^{48}$K, is in contradiction with $I^{\pi}=(1^{-})$ proposed by Kr\'{o}las \textit{et al.} \cite{Krolas2011}. In addition, the nuclear spin of the ground state of $^{50}$K was proposed to be $0^{-}$ \cite{Baumann1998} in contrast to the recent $\beta$-decay studies where it was suggested to be $1^{-}$ \cite{Crawford2009}. The ground state spin-parity of $^{49}$K was tentatively assigned to be ($1/2^+$) by Broda {\it{et al.}} \cite{Broda2010}, contrary to the earlier tentative $(3/2^+)$ assignment from $\beta$-decay spectroscopy \cite{Carraz1982}. For $^{51}$K, the nuclear spin was tentatively assigned to be $(3/2^{+})$ by Perrot \textit{et al.} \cite{Perrot2006}. Our recent hyperfine structure measurements of potassium isotopes using the collinear laser spectroscopy technique provided unambiguous spin values for $^{48-51}\rm{K}$ and gave the answer to the question as to what happens with the proton $sd$ orbitals for isotopes beyond $N = 28$. By measuring the nuclear spins of $^{49}\rm{K}$ and $^{51}\rm{K}$ to be $1/2$ and $3/2$ \cite{Papuga2013} respectively, the evolution of these two states in the potassium isotopes is firmly established. This is presented in Fig.\,\ref{fig: Levels-37-51K} for isotopes from $N=18$ up to $N=32$ where the inversion of the states is observed at $N=28$ followed by the reinversion back at $N=32$. \begin{figure} \includegraphics[width=1.0\linewidth]{Energy-37-51K} \caption{\label{fig: Levels-37-51K}(color online) Experimental energies for $1/2^{+}$ and $3/2^+$ states in odd-A K isotopes. Inversion of the nuclear spin is obtained in $^{47,49}$K and reinversion back in $^{51}$K. Results are taken from \cite{Measday2006,Huck1980,Weissman2004,Broda2010}. Ground-state spin for $^{49}$K and $^{51}$K were established \cite{Papuga2013}.} \end{figure} In addition, we have confirmed a spin-parity $1^{-}$ for $^{48}$K and $0^{-}$ for $^{50}$K \cite{Kreim2014}. The measured magnetic moments of $^{48-51}$K were not discussed in detail so far and will be presented in this article. Additionally, based on the comparison between experimental data and shell-model calculations, the configuration of the ground-state wave functions will be addressed as well. Finally, \textit{ab initio} Gorkov-Green's function calculations of the odd-A isotopes will be discussed. \section{Experimental procedure} The experiment was performed at the collinear laser spectroscopy beam line COLLAPS \cite{Mueller1983} at ISOLDE/CERN. The radioactive ion beam was produced by 1.4-GeV protons (beam current about 1.7\,$\mu$A) impinging on a thick UC$_{\rm{x}}$ target (45\,g/cm$^{2}$). Ionization of the resulting fragments was achieved by the surface ion source. The target and the ionizing tube were heated to around 2000\,$^{0}$C. The accelerated ions (up to 40\,kV) were mass separated by the high resolution separator (HRS). The gas-filled Paul trap (ISCOOL) \cite{Franberg2008,Vingerhoets2010} was used for cooling and bunching of the ions. Multiple bunches spaced by 90\,ms were generated after each proton pulse. The bunched ions were guided to the setup for collinear laser spectroscopy where they were superimposed with the laser. A schematic representation of the beam line for collinear laser spectroscopy is shown in Fig.~\ref{beam-line-COLLAPS}. \begin{figure} \includegraphics[width=1.0\linewidth]{beam-line} \caption{\label{beam-line-COLLAPS}(color online) Schematic representation of the setup for collinear laser spectroscopy at ISOLDE.} \end{figure} A cw titanium:sapphire (Ti:Sa) laser was operated close to the Doppler shifted $4s\;{^2S_{1/2}} \rightarrow 4p\;{^2P_{1/2}}$ transition at 769.9\,nm, providing around 1\,mW power into the beam line. Stabilization of the laser system during the experiment was ensured by locking the laser to a reference Fabry-Perot interferometer maintained under vacuum, which in turn was locked to a frequency stabilized helium-neon (HeNe) laser. An applied voltage of $\pm$10\,kV on the charge exchange cell (CEC) provided the Doppler tuning for the ions, which were neutralized through the collisions with potassium vapor. Scanning of the hfs was performed by applying an additional voltage in a range of $\pm$500\,V. The resonance photons were recorded by four photomultiplier tubes (PMT) placed immediately after the CEC. By gating the signal on the PMTs to the fluorescence photons from the bunches, the signal was only recorded for about 6\,$\mu$s when the bunches were in front of the PMTs. Consequently, the background related to the scattered laser light was suppressed by a factor $\sim 10^{4}$ (6\,$\mu$s/90\,ms). More details about the setup can be found in Ref.~\cite{Kreim2014}. \section{Results} In Fig.~\ref{Spectra} typical hyperfine spectra for $^{48-51}$K are shown. \begin{figure} \includegraphics[width=1.0\linewidth]{spectra-48-51K} \caption{\label{Spectra}(color online) The hyperfine spectra of $^{48-51}$K (a-d) obtained by collinear laser spectroscopy. The spectra are shown relative to the centroid of $^{39}$K.} \end{figure} The raw data are saved as counts versus scanning voltage. The conversion from voltage to frequency was carried out by using the masses from \cite{Wang2012} and applying the relativistic Doppler formula. The spectra were fitted with a Voigt line shape using common width for all components. The $\chi^2$-minimization procedure MINUIT \cite{James1994} was used with $A$-parameters ($A(S_{1/2})$ and $A(P_{1/2})$), the center of gravity and the intensities left as free fit parameters. Nuclear spins, magnetic moments and changes in mean square charge radii were extracted model independently. From the intensity ratios of the hyperfine components, the nuclear spin of $^{48}$K and $^{51}$K were determined to be $I =1$ \cite{Kreim2014} and $I =3/2$ \cite{Papuga2013}, respectively. Since only three peaks are observed in the hyperfine spectrum of $^{49}$K, a spin of $I = 1/2$ can be unambiguously assigned \cite{Papuga2013}. A single peak in the hyperfine spectrum of $^{50}$K corresponds to $I = 0$ \cite{Kreim2014}. The deduced magnetic moments and the implication for the nuclear structure of the potassium isotopes will be reported in this article. The observed hyperfine $A$-parameters of the ground and the excited states for all studied isotopes are presented in Table~\ref{Results}. \begin{table*} \caption{Magnetic hyperfine parameters for neutral potassium from this work and comparison with literature values \cite{Touchard1982,Phillips1965,Chan1969}. \label{Results}} \begin{ruledtabular} \begin{tabular}{c c x x x x } Isotope & \multicolumn{1}{c}{$I^{\pi}$} & \multicolumn{1}{c}{$A(^{2}S_{1/2})$ (MHz)} & \multicolumn{1}{c}{$A(^{2}P_{1/2})$ (MHz)} & \multicolumn{1}{c}{$A_{\rm{lit}}(^{2}S_{1/2})$ (MHz)} & \multicolumn{1}{c}{$A_{\rm{lit}}(^{2}P_{1/2})$ (MHz)} \\ \hline $^{38}$K & $3^{+}$ & +404.3\,(3) & +48.9\,(2) & +404.369\,(3) & - \\ $^{38 \rm m}$K & $0^{+}$ & 0 & 0 & - & - \\ $^{39}$K & $3/2^{+}$ & +231.0\,(3) & +27.8\,(2) & +231.0\,(3) & +27.5\,(4) \\ $^{42}$K & $2^{-}$ & -503.7\,(3) & -61.2\,(2) & -503.550779\,(5) & -60.6\,(16) \\ $^{44}$K & $2^{-}$ & -378.9\,(4) & -45.8\,(2) & -378.1\,(11) & -44.9\,(11) \\ $^{46}$K & $2^{-}$ & -462.8\,(3) & -55.9\,(2) & -465.1\,(12) & -55.7\,(13) \\ $^{47}$K & $1/2^{+}$ & +3413.2\,(3)\footnotemark[1] & +411.8\,(2) & +3420.2\,(29) & +411.9\,(50) \\ $^{48}$K & $1^{-}$ & -795.9\,(3) & -96.3\,(3) & - & - \\ $^{49}$K & $1/2^{+}$ & +2368.2\,(14) & +285.6\,(7) & - & - \\ $^{50}$K & $0^{-}$ & 0 & 0 & - & - \\ $^{51}$K & $3/2^{+}$ & +302.5\,(13) & +36.6\,(9) & - & - \\ \end{tabular} \end{ruledtabular} \footnotetext[1]{After reanalysis, the uncertainty on this value was increased from 0.2 to 0.3\,MHz.} \end{table*} The results are compared to the literature values from \cite{Phillips1965,Touchard1982,Chan1969}. Compared to the results from earlier atomic beam laser spectroscopy studies \cite{Touchard1982}, the precision has been increased by an order of magnitude for most of the values. The hyperfine $A$-parameters for $^{48-51}$K were measured for the first time. For the isotopes$/$isomer with $I = 0$, there is no hyperfine splitting of the atomic states, thus the $A$-parameters are equal to 0. The relation between $A$-parameters and magnetic moments is given by: $A = \mu B_{0} / I J$, where $B_{0}$ is the magnetic field induced by the electron cloud at the position of the nucleus. As $B_{0}$ is to first order isotope independent, magnetic moments were deduced relative to $^{39}$K using Eq.~(\ref{eq:Relative_mm}): \begin{equation} \label{eq:Relative_mm} \mu = \frac{A(^{2}S_{1/2})I}{A_{\rm{ref}}(^{2}S_{1/2})I_{\rm{ref}}}\mu_{\rm{ref}}. \end{equation} The reference values were taken from atomic-beam magnetic resonance measurements, where precise values are reported to be $A_{\rm{ref}}(^{2}S_{1/2})= + 230.8598601(7)$\,MHz and $\mu_{\rm{ref}}=+0.3914662(3)\,\mu_{N}$ \cite{Beckmann1974}. As the magnetic moments of potassium isotopes were determined with $10^{-3}$\,-\,$10^{-4}$ relative precision, one can not neglect the hyperfine structure (hfs) anomaly between two isotopes, arising from the finite size of the nuclei. This slightly modifies the $A$-parameters \cite{Bohr1951} and gives a small correction of Eq.~(\ref{eq:Relative_mm}) which is expressed by \begin {equation} ^{39}\Delta^{\rm{A}} = \frac{A^{39}(S_{1/2})/g(^{39}\rm{K})}{A^{\rm{A}}(S_{1/2})/g(^{A}\rm{K})} - 1 , \label{eq: Def-HFA} \end {equation} being different from zero. In Eq.~(\ref{eq: Def-HFA}), the $g$ factor is $g = \mu/I$. The dominant contributions to the hfs anomaly are originating from the difference in the nuclear magnetization distribution (Bohr-Weisskopff effect \cite{BohrWeisskopf1950}) and difference of the charge distribution (Breit-Rosenthal effect \cite{Rosenthal1932}). In the case of potassium isotopes, the hfs anomaly was measured for $^{38-42}$K relative to $^{39}$K \cite{Phillips1965,Eisinger1952,Beckmann1974,Chan1969,Ochs1950}. In order to assess the additional uncertainty on the magnetic moments for all measured isotopes, the hfs anomaly was estimated from the experimental data as well as from theoretical calculations. According to the approach proposed by Ehlers {\it{et al.}} \cite{Ehlers1968}, the differential hyperfine structure anomaly ($^{39}\delta^{\rm{A}}$) between two different electronic states is defined as: \begin {equation} ^{39}\delta^{\rm{A}} = \frac{A^{39}(S_{1/2})/A^{39}(P_{1/2})}{A^{\rm{A}}(S_{1/2})/A^{\rm{A}}(P_{1/2})} - 1 , \label{eq: Exp-HFA} \end {equation} where the $A$-parameters for the reference isotope $^{39}$K were taken from literature \cite{Beckmann1974, Falke2006}. The value of the hyperfine structure anomaly can be approximated by the differential hyperfine structure anomaly, which is good to a few percent. This is good enough considering the accuracies of our experimental results. Differential hyperfine anomalies are presented in Table~\ref{HF_anomaly} (col. 3). For $^{40,41}$K, the experimental results from literature were used: the $A(S_{1/2})$ parameter from \cite{Eisinger1952,Beckmann1974}, while the $A(P_{1/2})$ parameters were taken from \cite{Falke2006}. It should be noted that for $^{43,45}$K no data for $A(P_{1/2})$ were obtained. In addition, theoretical calculations were performed following Bohr \cite{Bohr1951}. The hfs anomaly was estimated to be $^{39}\Delta^{\rm{A}}_{\rm{theo}} = \epsilon (^{39}\rm{K})- \epsilon(^{\rm{A}}\rm{K})$, where $\epsilon (^A\rm{K})$ is a perturbation factor due to the finite size of the nucleus. It can be calculated using \cite{Bohr1951}: \begin {equation} \epsilon=-[(1+0.38\zeta)\alpha_{s}+0.62\alpha_{l}]b(Z, R_{0})(R/R_{0})^2. \label{eq: Epsilon-HFA} \end {equation} In Ref.~\cite{Bohr1951}, all parameters from Eq.~(\ref{eq: Epsilon-HFA}) are defined and for some of them values are tabulated. Theoretical estimations of the $\epsilon$ parameter and hfs anomaly ($^{39}\Delta^{\rm{A}}_{\rm{theo}}$) are listed in Table~\ref{HF_anomaly} (col. 4 and 5). Hyperfine structure anomalies of the potassium isotopes known from the literature \cite{Ochs1950,Beckmann1974,Eisinger1952,Chan1969,Phillips1965} are shown in the last column of Table~\ref{HF_anomaly} ($^{39}\Delta_{\rm{lit}}^{\rm{A}}$). \begin{table} \caption {Estimated hyperfine structure anomalies of potassium isotopes. Experimental results for the hyperfine parameters were used to calculate ($^{39}\delta^{\rm{A}}$) from Eq.\,(\ref{eq: Exp-HFA}). For $^{40,41}$K experimental data were taken from \cite{Eisinger1952,Beckmann1974,Falke2006}. The $\epsilon (^A\rm{K})$ parameters for all isotopes are calculated from Eq.\,(\ref{eq: Epsilon-HFA}) and are listed in the next column. For the reference isotope, it was found to be $\epsilon (^{39}\rm{K}) = 0.165$. The estimated hyperfine structure anomalies from the model ($^{39}\Delta_{\rm{theo}}^{\rm{A}}$) described by Bohr (see text for details) are shown as well. In the last column, the hyperfine structure anomalies from literature ($^{39}\Delta_{\rm{lit}}^{\rm{A}}$) are given \cite{Ochs1950,Beckmann1974,Eisinger1952,Chan1969,Phillips1965}. \label{HF_anomaly}} \begin{ruledtabular} \begin{tabular}{c c c c c c} Isotope & \multicolumn{1}{c}{$I^{\pi}$} & \multicolumn{1}{c}{$^{39}\delta^{\rm{A}}$ (\%)} & \multicolumn{1}{c}{$\epsilon (^A\rm{K})$} & \multicolumn{1}{c}{$^{39}\Delta_{\rm{theo}}^{\rm{A}}$ (\%)} & \multicolumn{1}{c}{$^{39}\Delta_{\rm{lit}}^{\rm{A}}$ (\%)} \\ \hline $^{38}$K & $3^{+}$ & 0.53\,(44) & -0.006 & 0.17 & 0.17\,(6) \\ $^{40}$K & $4^{-}$ & 0.43\,(17) & -0.379 & 0.54 & 0.466\,(19) \\ $^{41}$K & $3/2^{+}$ & -0.23\,(31) & 0.398 & -0.23 & -0.226\,(10) \\ & & & & & -0.22936\,(14) \\ $^{42}$K & $2^{-}$ & 0.99\,(36) & -0.265 & 0.43 & 0.336\,(38) \\ $^{43}$K & $3/2^{+}$ & - & 0.560 & -0.39 & -\\ $^{44}$K & $2^{-}$ & 0.47\,(47) & -0.302 & 0.47 & - \\ $^{45}$K & $3/2^{+}$ & - & 0.521 & -0.36 & - \\ $^{46}$K & $2^{-}$ & 0.40\,(39) & -0.275 & 0.44 & - \\ $^{47}$K & $1/2^{+}$ & 0.28\,(16) & -0.126 & 0.29 & - \\ $^{48}$K & $1^{-}$ & 0.57\,(35) & -0.211 & 0.38 & - \\ $^{49}$K & $1/2^{+}$ & 0.24\,(29) & -0.121 & 0.29 & - \\ $^{51}$K & $3/2^{+}$ & 0.57\,(250) & 0.097 & 0.07 & - \\ \end{tabular} \end{ruledtabular} \end{table} For all isotopes except $^{42}$K, the hyperfine structure anomaly estimated from the experimental results is in agreement with the calculated ones. The values for odd-odd nuclei are systematically higher than for odd-even, thus we will quote different additional uncertainties on the magnetic moments (in square brackets in Table~\ref{mm-odd-A-Tab} and Table~\ref{mm-even-A}), namely 0.3\% and 0.5\% for odd-A and even-A isotopes, respectively. \section{Discussion} Nuclei with one particle or one hole next to a shell closure are excellent probes for testing shell-model interactions. In this context, the investigation of the potassium chain is of great interest, since it has a hole in the $\pi sd$ orbital and it covers two major neutron shells, $N = 20$ and $N = 28$, and one sub-shell at $N = 32$. In what follows, the experimental results from our work are compared to shell-model predictions. The calculations were carried out using the ANTOINE code \cite{Nowacki1999} for two effective interactions: SDPF-NR \cite{Retamosa1997, Nummela2001-ANT} and SDPF-U \cite{Nowacki2009}. The latter is a more recent version of the SDPF-NR interaction where the monopole part was refitted by including more experimental results from nuclei with one particle or one hole next to the closed shell for protons or neutrons such as $^{35}$Si, $^{47}$Ar and $^{41}$Ca. The calculations have been performed in the $0\hbar \omega$ shell model space beyond a $^{16}$O core and with valence protons restricted to $sd$ orbitals and neutrons to $sd$ or $pf$ orbitals. Neutron excitations across $N = 20$ were prohibited. In order to account for missing interactions among the valence nucleons as well as with the nucleons from the core, the calculations were performed using effective $g$ factors: the spin $g$ factors were fixed at $g_{s}^{\rm{eff}}=0.85 g_{s}^{\rm{free}}$, while the orbital $g$ factors were fixed to $g_{l}^{\pi} = 1.15$ and $g_{l}^{\nu}$ = -0.15 \cite{Richter2008}. \subsection{Odd-A} Nuclear properties such as the ground-state spin and magnetic moment of odd-A K isotopes (odd-even isotopes) are determined by an unpaired proton placed in the $\pi sd$ orbital whilst the even number of neutrons are coupled to spin zero. In the simple shell-model framework the measured nuclear spin indicates the dominant component of the ground-state wave function. Based on this simple model, one would expect that the magnetic moments of these isotopes are equal to the single-particle magnetic moments of the orbital where a valence proton is located. However, the observed deviation from the single-particle values reveals influence of the proton-neutron interaction leading to a more collective behavior. Although the magnetic moments of the neutron-rich odd-A K isotopes were already published in \cite{Papuga2013}, a detailed discussion over the entire odd-A chain from $N = 20$ up to $N = 32$ will be presented here with additional focus on the monopole interaction responsible for the shell evolution. The experimentally observed magnetic moments are listed in Table~\ref{mm-odd-A-Tab} together with the values predicted by shell-model calculations using the SDPF-NR and SDPF-U effective interactions. \begin{table*} \caption{Experimental magnetic moments (in units of $\mu_{\rm{N}}$) compared with the calculated ones using two effective interactions: SDPF-NR and SDPF-U. The predicted amount of the $\pi 1d_{3/2}^{-1}\otimes \nu (fp)$ in the ground-state wave function is given in \%. If available, the literature values are shown as well. The uncertainty in the square brackets is due to the hyperfine structure anomaly and is 0.3\%. \label{mm-odd-A-Tab}} \begin{ruledtabular} \begin{tabular}{c c a c c c c z c} Isotope & \multicolumn{1}{c}{$I^{\pi}$} & \multicolumn{1}{c}{$\mu_{\rm exp}$} & \multicolumn{1}{c}{$\mu_{\rm{SDPF-NR}}$} & \multicolumn{1}{c}{$\pi 1d_{3/2}^{-1}$ (\%)} & \multicolumn{1}{c}{$\mu_{\rm{SDPF-U}}$} & \multicolumn{1}{c}{$\pi 1d_{3/2}^{-1}$ (\%)} & \multicolumn{1}{c}{$\mu_{\rm lit}$} & \multicolumn{1}{c}{Reference} \\ \hline $^{39}$K & $3/2^{+}$ & +0.3917\,(5)\,[12] & +0.65 & 100\% & +0.65 & 100\% & +0.3914662\,(3) & \cite{Beckmann1974} \\ $^{41}$K & $3/2^{+}$ & - & +0.37 & 95\% & +0.33 & 95\% & +0.2148701\,(2) & \cite{Beckmann1974} \\ $^{43}$K & $3/2^{+}$ & - & +0.22 & 92\% & +0.17 & 92\% & +0.1633\,(8)\footnotemark[1] & \cite{Touchard1982} \\ $^{45}$K & $3/2^{+}$ & - & +0.23 & 88\% & +0.21 & 90\% & +0.1734\,(8)\footnotemark[1] & \cite{Touchard1982} \\ $^{47}$K & $1/2^{+}$ & +1.9292\,(2)\,[58] & +1.87 & 13\% & +1.91 & 13\% & +1.933(9)\footnotemark[1] & \cite{Touchard1982} \\ $^{49}$K & $1/2^{+}$ & +1.3386\,(8)\,[40] & +1.61 & 21\% & +1.81 & 15\% & - & - \\ $^{51}$K & $3/2^{+}$ & +0.5129\,(22)\,[15] & +0.60 & 90\% & +0.65 & 93\% & - & - \\ \end{tabular} \end{ruledtabular} \footnotetext[1]{Included 0.5\% uncertainty on the error to account for the hyperfine structure anomaly.} \end{table*} In the same table, the calculated percentage of the component of the ground-state wave function originating from a hole in the $\pi 1d_{3/2}^{-1}$ is shown as well. In Fig.~\ref{Fig-mm-odd-A} the experimental magnetic moments for odd-A K isotopes are compared to the results from the shell-model calculations. \begin{figure} \includegraphics[width=1.0\linewidth]{mm-odd-A-eff} \caption{\label{Fig-mm-odd-A} (color online) Experimental magnetic moments (black dots) compared to the shell-model calculation using SDPF-NR (red dashed line) and SDPF-U (blue solid line) interactions and effective $g$ factors (see text for more details). In general a very good agreement between experimental and theoretical results is observed, except for $^{39}$K and $^{49}$K.} \end{figure} In general a very good agreement between experimental and theoretical results is observed. The discrepancy for $^{39}$K and $^{41}$K might be due to excitation across the $Z,N=20$ shell gaps, which were not considered in these calculations. This problem is especially pronounced for $^{39}$K where, the shell-model calculations yield a pure $\pi 1d_{3/2}^{-1}$ state with a magnetic moment about $60$\% larger than the experimental value. Both effective interactions yield almost identical amounts of the $\pi 1d_{3/2}^{-1}\otimes \nu (pf)$ component in the ground state of odd-A isotopes. It is more than 90\% for all isotopes up to $^{45}$K, but for $^{47,49}$K the wave function is dominated by the $\pi 2s_{1/2}^{-1} \otimes \nu (pf)$ configuration. This was already concluded in \cite{Papuga2013}, the conclusion based on the measured ground-state spin and $g$ factor. The only noticeable difference between both calculations is found for $^{49}$K, where the contribution from $\pi 1d_{3/2}^{-1} \otimes \nu (pf)$ is predicted to be 21\% from SDPF-NR and 15\% from SDPF-U. In both cases the calculated value deviates from the experimental one, but SDPF-U shows a larger deviation. From a two-state mixing calculation, at least 25\,\% \cite{Neyens2013} of mixing with the $[\pi 1d_{3/2}^{-1} \otimes \nu (pf)_{2^+}]_{1/2^{+}}$ is needed to reproduce the observed magnetic moment. The inversion of the nuclear spin from $I = 3/2$ to $I = 1/2$ at $N = 28$ and the re-inversion back to $I=3/2$ at $N=32$ is related to the evolution of the proton orbitals ($\pi sd$) while different neutron orbitals are being filled. This evolution is driven by the monopole term of the nucleon-nucleon ($NN$) interaction. According to Otsuka {\it{et al.}} \cite{Otsuka2010}, the interaction has a linear dependence on the occupation number and consists of three parts: central, vector and tensor. Applying the spin-tensor decomposition method \cite{Smirnova2010,Smirnova2012}, it is possible to separate the contribution of different components of the effective $NN$ interaction. This leads to a qualitative analysis of the role of each part separately in the evolution of the effective single-particle energies (ESPEs). The calculated centroids for every component of the monopole interaction are listed in Table\,\ref{tab: Centroids-SDPF-NR-U}. \begin{table*} \centering \caption {Spin-tensor content of the centroids of the SDPF-NR and SDPF-U interaction, defining the proton $1d_{3/2}-2s_{1/2}$ gap. Results are presented in MeV.\label{tab: Centroids-SDPF-NR-U}} \begin{ruledtabular} \begin{tabular}{c c x x x x x x} Interaction & Component & \multicolumn{1}{c}{$V_{d_{3/2}f_{7/2}}^{\pi \nu}$} & \multicolumn{1}{c}{$V_{s_{1/2}f_{7/2}}^{\pi \nu}$} & \multicolumn{1}{c}{$\Delta V$} & \multicolumn{1}{c}{$V_{d_{3/2}p_{3/2}}^{\pi \nu}$} & \multicolumn{1}{c}{$V_{s_{1/2}p_{3/2}}^{\pi \nu}$} & \multicolumn{1}{c}{$\Delta V$} \\ \hline \space & Central & -1.66 & -1.26 & -0.40 & -1.34 & -1.46 & +0.12\\ SDPF-NR & Vector & +0.28 & +0.17 & +0.11 & +0.21 & +0.22 & -0.01\\ \space & Tensor & -0.28 & 0.00 & -0.28 & -0.08 & 0.00 & -0.08\\ \cline{2-8} \space & Total & -1.66 & -1.09 & -0.57 & -1.21 & -1.24 & +0.03 \\ \hline \space & Central & -1.51 & -1.21 & -0.30 & -1.05 & -1.21 & +0.16 \\ SDPF-U & Vector & +0.09 & +0.07 & +0.02 & +0.05 & -0.11 & +0.16 \\ \space & Tensor & -0.28 & 0.00 & -0.28 & -0.06 & 0.00 & -0.06\\ \cline{2-8} \space & Total & -1.70 & -1.14 & -0.56 & -1.06 & -1.32 & +0.26 \\ \end{tabular} \end{ruledtabular} \end{table*} The centroid of the proton-neutron interaction is defined as \cite{Poves1981}: \begin{equation} V_{j_\pi j_\nu}=\frac{\sum_{J}(2J+1){\langle j_\pi j_\nu|V|j_\pi j_\nu\rangle}}{\sum_{J}(2J+1)}, \end{equation} where $j_\pi$ and $j_\nu$ stand for the angular momentum of proton and neutron orbitals, $\langle j_\pi j_\nu|V|j_\pi j_\nu \rangle$ is the two-body matrix element and $J$ is the total angular momentum of a proton-neutron state. The summation runs over all possible values of $J$. Based on the results presented in Table\,\ref{tab: Centroids-SDPF-NR-U}, the central component of the interaction is by far the largest (col. 3-4 and 6-7) and, thus has the strongest influence on the energy shift. Note that there is no tensor component for the $s_{1/2}$ orbital due to the absence of a preferred orientation of the spin for an $l = 0$ state \cite{Otsuka2005}. The change of the energy gap between $\pi 1d_{3/2}$ and $\pi 2s_{1/2}$ depends on the difference $\Delta V$ between the two centroids (Table\,\ref{tab: Centroids-SDPF-NR-U}; col. 5 and 8). The evolution of the energy gap from $N=20$ to $N=28$ and from $N=28$ to $N=32$, along with the spin-tensor decomposition of this energy gap, is presented in Table\,\ref{tab: ST-decomposition}. \begin{table} \caption {Calculated contributions of the different spin-tensor terms of SDPF-NR ("NR") and SDPF-U ("U") to the evolution of the energy gap between effective $\pi 1d_{3/2}$ and $\pi 2s_{1/2}$ when filling $\nu 1f_{7/2}$ and $\nu 2p_{3/2}$ orbitals. The results are given in MeV. \label{tab: ST-decomposition}} \begin{ruledtabular} \begin{tabular}{c c c c c} filling & \multicolumn{2}{c}{$\nu 1f_{7/2}$} & \multicolumn{2}{c}{$ \nu 2p_{3/2}$} \\ \hline \space & \rm{NR} & \rm{U} & \rm{NR} & \rm{U} \\ \hline Central & -2.09 & -1.58 & +0.46 & +0.58 \\ Vector & +0.58 & +0.06 & -0.06 & +0.43 \\ Tensor & -1.64 & -1.64 & -0.17 & -0.12 \\ \hline Total & -3.15 & -3.16 & +0.23 & +0.89 \\ \end{tabular} \end{ruledtabular} \end{table} Both interactions predict the same decrease of the gap by $-3.15\,\rm{MeV}$ for isotopes from $N = 20$ up to $N = 28$ (Table\,\ref{tab: ST-decomposition}; col. 2-3), although the central and vector contribution are significantly different in both interactions. Once the $\nu p_{3/2}$ orbital is involved, for isotopes from $N=29$ up to $N=32$, the situation changes. The increase in the gap between $\pi 1d_{3/2}$ and $\pi 2s_{1/2}$ (Table\,\ref{tab: ST-decomposition}; col. 4-5) is mostly driven by the central component in the SDPF-NR interaction, while also the vector component contributes significantly in the SDPF-U. Therefore, the calculated change in the energy gap is very different: +0.23\,MeV and +0.89\,MeV, respectively. This results in different calculated spectra for $^{49}$K and $^{51}$K as illustrated in Fig.~\ref{Energy-diff}. This figure shows the energy difference between the lowest $1/2^{+}$ and $3/2^{+}$ states for isotopes in the range from $N = 24$ up to $N = 34$ compared to the calculated values. \begin{figure}[h!] \includegraphics[width=0.9\linewidth]{Energy-diff-s-d} \caption{\label{Energy-diff} (color online) Energy difference between the two lowest states with $I^{\pi}=1/2^{+}$ and $3/2^{+}$ for odd-A K isotopes from $N = 24$ up to $N = 34$. Experimental results (black stars) taken from \cite{Measday2006,Huck1980,Weissman2004,Broda2010} are in good agreement with the shell-model calculations using different effective interactions: SDPF-NR (red dots) and SDPF-U (blue triangles). For $^{49}$K, only the SDPF-NR interaction correctly predicts the spin of the ground state to be $1/2^{+}$. The shaded area represents the expected region based on the measured ground-state spin and the shell-model calculation for the first excited state in $^{51}$K. } \end{figure} Up to $N= 28$, both interactions are in agreement with the experimental results. The deviation between both effective interactions increases beyond $N = 28$ when the $\nu 2p_{3/2}$ and higher orbitals are involved. For $^{49}$K, both interactions calculate the energy difference between the ground and first excited state to be about 75\,keV, but only the SDPF-NR predicts the correct ground-state spin. Although both effective interactions predicted the correct ground-state spin for $^{51}$K, experimental data on the energy of the first-excited state is needed to further test the validity of both models. Beyond $N=32$ the predicted ground-state spin 3/2 for $^{53}$K needs experimental verification, as well as the energy of the first excited $1/2^+$ state, which is very different in both calculations. Very recently, {\it{ab initio}} calculations of open-shell nuclei have become possible in the Ca region~\cite{Soma2014} on the basis of the self-consistent Gorkov-Green`s function formalism~\cite{Soma2011}. State-of-the-art chiral two- ($NN$)~\cite{Entem2003,Machleidt2011} and three-nucleon ($3N$)~\cite{Navratil2007} interactions adjusted to two-, three- and four-body observables (up to $^4$He) are employed, without any further modification, in the computation of systems containing several tens of nucleons. We refer to Ref.~\cite{Soma2014} for further details. In the present study, Gorkov-Green's function calculations of the lowest $1/2^{+}$ and $3/2^{+}$ states in $^{37-53}$K have been performed by removing a proton from $^{38-54}$Ca. Similarly to Fig.~\ref{Energy-diff}, the upper panel of Fig.~\ref{AI-ESPE} compares the results to experimental data. The calculated energy differences have been shifted down by 2.58 MeV to match the experimental value for $^{47}$K. The overestimation of energy differences is a general feature of calculated odd-A spectra and actually correlates with the systematic overbinding of neighboring even-A ground states \cite{Soma2014}. Still, one observes the correct {\it relative} evolution of the $1/2^{+}$ state with respect to the $3/2^{+}$ when going from $^{37}$K to $^{47}$K and then from $^{47}$K to $^{49}$K. This result is very encouraging for these first-ever systematic {\it{ab initio}} calculations in mid-mass nuclei. Indeed, it allows one to speculate that correcting in the near future for the systematic overbinding produced in the Ca region by currently available chiral interactions, and thus for the too spread out spectra of odd-A systems, might bring the theoretical calculation in good agreement with experiment. Although this remains to be confirmed, it demonstrates that systematic spectroscopic data in mid-mass neutron-rich nuclei provide a good test case to validate/invalidate specific features of basic inter-nucleon interactions and innovative many-body theories. \begin{figure}[h!] \includegraphics[width=0.9\linewidth]{ab-initio-final} \caption{\label{AI-ESPE} (color online) (a) Energy difference between the lowest $1/2^{+}$ and $3/2^{+}$ states obtained in $^{37-53}$K from {\it{ab initio}} Gorkov-Green`s function calculations and experiment. {\it{Ab initio}} results have been shifted by 2.58\,MeV to match the experimental $(1/2^{+}-3/2^{+})$ splitting in $^{47}$K. (b) $\pi d_{3/2}$ and $\pi s_{1/2}$ effective single-particle energies (ESPE) in $^{37-53}$K calculated in Gorkov-Green's functions theory. } \end{figure} To complement the above analysis, the lower panel of Fig.~\ref{AI-ESPE} provides the evolution of proton $1d_{3/2}$ and $2s_{1/2}$ shells. These two effective single-particle energies (ESPEs) recollect \cite{Soma2011} the fragmented $3/2^{+}$ and $1/2^{+}$ strengths obtained from one-proton addition and removal processes on neighboring Ca isotones. Within the present theoretical description, the evolution of the observable (i.e. theoretical-scheme independent) lowest-lying $1/2^{+}$ and $3/2^{+}$ states does qualitatively reflect the evolution of the underlying non-observable (i.e. theoretical-scheme dependent) single-particle shells. As such, the energy gap between the two shells decreases from 5.76\,MeV in $^{39}$K to 1.81\,MeV in $^{47}$K which is a reduction of about $70$\%. Adding 4 neutrons in the $\nu 2p_{3/2}$ shell causes the energy difference to increase again to 4.03\,MeV. \subsection{Even-A} The configuration of the even-A potassium isotopes arises from the coupling between an unpaired proton in the $sd$ shell with an unpaired neutron. Different neutron orbitals are involved, starting from $^{38}$K where a hole in the $\nu 1d_{3/2}$ is expected, then gradually filling the $\nu 1 f_{7/2}$ and finally, the $\nu 2p_{3/2}$ for $^{48,50}$K. In order to investigate the composition of the ground-state wave functions of the even-A K isotopes, we first compare the experimental magnetic moments to the semi-empirical values. Based on the additivity rule for the magnetic moments ($g$ factors) and assuming a weak coupling between the odd proton and the odd neutron, the semi-empirical magnetic moments can be calculated using the following formula \cite{Schwartz1953}: $\mu_{\rm{se}}=g_{\rm{se}}\cdot I$, with \begin{equation} \label{eq:Empirical_g-factors} \resizebox{.89\hsize}{!}{$g_{\rm{se}}=\frac{g(j_{\pi})+g(j_{\nu})}{2}+\frac{g(j_{\pi})-g(j_{\nu})}{2}\frac{j_{\pi}(j_{\pi}+1)-j_{\nu}(j_{\nu}+1)}{I(I+1)}$}, \end{equation} where $g(j_{\pi})$ and $g(j_{\nu})$ are the experimental $g$ factors of nuclei with an odd proton or neutron in the corresponding orbital. The calculations were performed using the measured $g$ factors of the neighboring isotopes with the odd-even and even-odd number of particles in $j_{\pi}$ and $j_{\nu}$, respectively. For the empirical values of unpaired protons, results from Table\,\ref{mm-odd-A-Tab} were used. The $g$ factors for the odd neutrons were taken from the corresponding Ca isotones \cite{Minamisono1976,Andl1982,Olschewski1972,GarciaRuiz2013}. The obtained results with the list of isotopes used for different configurations are presented in Table~\ref{empirical_mm}. \begin{table}[h!] \caption {Semi-empirical $g$ factors obtained for certain configurations using the additivity rule in Eq.\,(\ref{eq:Empirical_g-factors}) (see text for more details). In the calculations, results from Table\,\ref{mm-odd-A-Tab} were used for $g(j_{\pi})$, while for $g(j_{\nu})$ Ca data were taken from \cite{Minamisono1976,Andl1982,Olschewski1972,GarciaRuiz2013}. For $^{48}$K, different configurations are considered for $I=1$. \label{empirical_mm}} \begin{ruledtabular} \begin{tabular}{c c c e c} Isotope & $I^{\pi}$ & configuration & \multicolumn{1}{c}{$g_{\rm{se}}$} & ($g(j_{\pi})$;$g(j_{\nu})$) \\ \hline $^{38}$K & $3^{+}$ & $\pi 1d_{3/2}^{-1} \otimes \nu 1d_{3/2}^{-1}$ & +0.47 & ($^{39}$K; $^{39}$Ca ) \\ $^{40}$K & $4^{-}$ & $\pi 1d_{3/2}^{-1} \otimes \nu 1f_{7/2}$ & -0.31 & ($^{39}$K; $^{41}$Ca) \\ $^{42}$K & $2^{-}$ & $\pi 1d_{3/2}^{-1} \otimes \nu 1f_{7/2}^{3}$ & -0.64 & ($^{41}$K; $^{43}$Ca) \\ $^{44}$K & $2^{-}$ & $\pi 1d_{3/2}^{-1} \otimes \nu 1f_{7/2}^{5}$ & -0.62 & ($^{43}$K; $^{45}$Ca) \\ $^{46}$K & $2^{-}$ & $\pi 1d_{3/2}^{-1} \otimes \nu 1f_{7/2}^{-1}$ & -0.65 & ($^{45}$K; $^{47}$Ca) \\ $^{48}$K & $1^{-}$ & $\pi 1d_{3/2}^{-1} \otimes \nu 2p_{3/2}$ & -0.40 & ($^{45}$K; $^{49}$Ca) \\ $^{48}$K & $1^{-}$ & $\pi 2s_{1/2}^{-1} \otimes \nu 2p_{3/2}$ & -2.11 & ($^{47}$K; $^{49}$Ca) \\ \end{tabular} \end{ruledtabular} \end{table} A comparison between the experimental and semi-empirical $g$ factors is shown in Fig.~\ref{Even-A-emp}. \begin{figure}[h!] \includegraphics[width=0.95\linewidth]{Empirical-corr} \caption{\label{Even-A-emp} (color online) Experimental $g$ factors (black dots) compared to the semi-empirical values (red solid line) calculated from the neighboring isotopes. Based on the good agreement between the experimental and semi-empirical $g$ factors, the dominant component of the wave functions can be easily established for $^{38-46}$K. Only for $^{48}$K a strong mixing between the $\pi 2s_{1/2}^{-1} \otimes \nu 2p_{3/2}$ and the $\pi 1d_{3/2}^{-1} \otimes \nu 2p_{3/2}$ in the wave function is found.} \end{figure} For $^{38}$K, the semi-empirical value calculated from $^{39}$K and $^{39}$Ca provides excellent agreement with the experimental result. This confirms that the dominant component in the wave function for the ground state originates from the coupling between a hole in the $\pi 1d_{3/2}$ and the $\nu 1d_{3/2}$. By adding more neutrons, the $\nu 1f_{7/2}$ orbital is filled for $^{40}$K up to $^{46}$K. In order to calculate the semi-empirical $g$ factors for these isotopes, $g(j_{\pi})$ is provided from neighboring odd-A K isotopes (Table\,\ref{mm-odd-A-Tab}) combined with $g(j_\nu)$ of the subsequent odd-A Ca isotones starting from $N = 21$ up to $N = 27$. The trend of the experimental $g$ factors is very well reproduced by the semi-empirical calculations suggesting that the dominant component in the wave function of these isotopes is $\pi 1d_{3/2}^{-1} \otimes \nu 1f_{7/2}^{n}$ where $n=1,3,5,7$. For $^{48}$K, two semi-empirical values are calculated by considering a coupling between a proton hole in the $\pi 2s_{1/2}$ or the $\pi 1d_{3/2}$ with neutrons in the $\nu 2p_{3/2}$ orbital. Comparing the experimental $g$ factor to the semi-empirical results, it is possible to conclude that the main component in the wave function of this isotope arises from the configuration with a hole in the $\pi 1d_{3/2}$. Nevertheless, the deviation of the experimental result from the semi-empirical $g$ factors is due to a large amount of mixing between both configurations in the wave function. $^{50}$K is not presented because the observed $I = 0$ leads to $\mu = 0$. There are two possible configurations which would yield this particular spin: $\pi 1d_{3/2}^{-1} \otimes \nu 2p_{3/2}$ and $\pi 2s_{1/2}^{-1} \otimes \nu 2p_{1/2}$. The experimental magnetic moments together with shell-model calculations are summarized in Table~\ref{mm-even-A} \begin{table*} \caption{Experimental magnetic moments (in units of $\mu_{\rm{N}}$) for even-A K isotopes compared to shell-model predictions using two effective interactions: SDPF-NR and SDPF-U. The error in the square brackets is due to the hyperfine structure anomaly, which amounts to 0.5\%. \label{mm-even-A}} \begin{ruledtabular} \begin{tabular}{c c z x x f c} Isotope & $I^{\pi}$ & \multicolumn{1}{c}{$\mu_{\rm exp}$} & \multicolumn{1}{c}{$\mu_{\rm{SDPF-NR}}$} & \multicolumn{1}{c}{$\mu_{\rm{SDPF-U}}$} & \multicolumn{1}{c}{$\mu_{\rm lit}$} & Reference \\ \hline $^{38}$K & $3^{+}$ & +1.3711\,(10)\,[69] & +1.33 & +1.33 & +1.371\,(6)\footnotemark[1] & \cite{Touchard1982} \\ $^{40}$K & $4^{-}$ & - & -1.63 & -1.63 & -1.2964\,(4)\footnotemark[2] & \cite{Eisinger1952}\\ $^{42}$K & $2^{-}$ & -1.1388\,(7)\,[57] & -1.58 & -1.56 & -1.14087\,(20)\footnotemark[2] & \cite{Chan1969}\\ $^{44}$K & $2^{-}$ & -0.8567\,(9)\,[43] & -1.05 & -0.90 & -0.856\,(4)\footnotemark[1] & \cite{Touchard1982}\\ $^{46}$K & $2^{-}$ & -1.0464\,(7)\,[52] & -1.21 & -1.18 & -1.051\,(6)\footnotemark[1] & \cite{Touchard1982}\\ $^{48}$K & $1^{-}$ & -0.8997\,(3)\,[45] & -0.77 & -0.55 & - & - \\ \end{tabular} \end{ruledtabular} \footnotetext[1]{Included 0.5\% uncertainty on the error to account for the hyperfine structure anomaly.} \footnotetext[2]{The value without diamagnetic correction of +0.13\%. } \end{table*} and graphically presented in Fig.~\ref{Even-A_mm_eff}. \begin{figure} \includegraphics[width=0.95\linewidth]{mm-even-A-eff} \caption{\label{Even-A_mm_eff}(color online) Measured magnetic moments (black dots) for even-A K isotopes compared to the shell-model calculations using the SDPF-NR (red dashed line) as well as the SDPF-U (blue solid line) effective interaction. Although there is a larger deviation present for $^{40}$K and $^{42}$K, which might originate from lack of the excitations across $Z, N =20$, overall reasonable agreement between the experimental and theoretical results is observed. } \end{figure} The predictions for $^{38}$K from both interactions reproduce the experimental magnetic moments very well. Furthermore, almost the same value is calculated with both interactions for $A=40$ and $A=42$, but the experimental results are underestimated by about 26\% and 37\% for $^{40}$K and $^{42}$K, respectively. While the SDPF-U interaction almost reproduces the observed magnetic moment for $^{44}$K, its earlier version (SDPF-NR) shows a deviation of approximately 0.26\,$\mu_{\rm{N}}$ when comparing to the experimental result. Additionally, very good agreement is observed between experimental and theoretical results for $^{46}$K. Finally, the situation is inverted for the case with the strongly mixed $^{48}$K, which is better reproduced by the SDPF-NR interaction and shows a deviation of about 0.35\,$\mu_{\rm{N}}$ for SDPF-U. The general trend of the magnetic moments is well reproduced by both interactions and the calculated magnetic moments are in reasonable agreement with the experimental results. The slightly larger deviation observed for $^{40}$K and $^{42}$K is probably due to lack of excitations across $Z, N = 20$. At this point one should be aware that the odd-odd isotopes are more challenging for the shell-model calculation than odd-even nuclei due to the high level density at low energy. These levels arise from all different possibilities of couplings between an odd proton and an odd neutron. Although the energy of a calculated level might be wrong by hundreds of keV, if the magnetic moment is well-reproduced we can still draw reliable conclusions on the wave function composition of the state. In the case of $^{38}$K, the $\pi 1d_{3/2}^{-1} \otimes \nu 1d_{3/2}^{-1}$ configuration constitutes more than 90\% of the total wave function. The dominant component of the ground-state wave function for all $N>20$ even-A K isotopes is arising from a hole in the $\pi 1d_{3/2}$ coupled to an odd neutron in the $pf$ orbital. For $^{40,42,44,46}$K, the main component is $\pi 1d_{3/2}^{-1} \otimes \nu 1f_{7/2}^{n}$ and its contribution to the wave function decreases from more than 90\% down to about 85\%. The lowest $1^-$ state in $^{48}$K is predicted to be an excited state by both interactions, respectively at $E = 407$\,keV and $E = 395$\,keV (see Fig.~\ref{EL-48K}). \begin{figure}[h!] \includegraphics[width=0.65\linewidth]{energy-levels-48K} \caption{\label{EL-48K} Experimental energy spectrum of $^{48}$K adapted from Ref.\,\cite{Krolas2011} using the fact that the nuclear spin is firmly established to be $1^{-}$ \cite{Kreim2014}. Results are compared to the calculated spectra from different effective interactions: SDPF-NR and SDPF-U. } \end{figure} Both interactions favor a $2^{-}$ state as the ground state. In addition, an excited $2^-$ state is near-degenerate with the $1^-$ level, at $E=408$\,keV and $E=340$\,keV respectively. Considering the firmly assigned ground-state spin-parity of $^{48}$K, and using the multipolarities deduced from the measured lifetimes of the lowest four levels by Kr\'{o}las \textit{et al.} \cite{Krolas2011}, the experimental spin-parities of the four lowest excited states can now be more firmly assigned. A reasonable agreement with the calculated level scheme is shown up to 1\,MeV. However, the positive parity level around 2\,MeV, which must be due to a proton excitation across the $Z=20$ gap, is not reproduced in the current calculations, as such excitations have not been included. The wave function of the calculated lowest $1^-$ state, which reproduces the observed magnetic moment reasonably well, is very fragmented compared to the other even-A K isotopes: $\pi 1d_{3/2}^{-1} \otimes \nu 2p_{3/2}$ only constitutes approximately 40\% and 50\% of the total wave function for SDPF-NR and SDPF-U, respectively. The next leading component, $\pi 2s_{1/2}^{-1} \otimes \nu 2p_{3/2}$, contributes only 15-20\%, although this isotope is located between two isotopes with a dominant $\pi 2s_{1/2}^{-1}$ configuration ($^{47}$K and $^{49}$K). In addition, configurations which arise from $1p1h$ excitation from $ \nu 1f_{7/2}$ to the rest of the $\nu (pf)$ shell have a significant contribution of about 15\% to the total wave function of the lowest $1^-$ state in $^{48}$K. In the case of $^{50}$K, the wave function of the $0^-$ level is much less fragmented: the main component is $\pi 1d_{3/2}^{-1} \otimes \nu (pf)$, constituting more than 85\% of the wave function. The contribution of the $\pi 2s_{1/2}^{-1}\otimes \nu (pf)$ component as well as the one from $1p1h$ neutron excitations is about 5\%. While this $0^-$ is correctly reproduced as the ground state by the SDPF-U interaction, it is predicted at 315\,keV (with a $2^-$ ground state) with SDPF-NR. In addition to the magnetic moment and wave functions obtained from the shell-model calculations, it is also possible to extract information about the occupancy of the orbitals. The calculated occupancy of the $\pi 2s_{1/2}$ and $\pi 1d_{3/2}$ orbitals are shown in Fig.~\ref{Occupation}. \begin{figure}[h!] \includegraphics[width=0.95\linewidth]{Occupation-NR-U-all} \caption{\label{Occupation}(color online) Proton occupation of the $\pi 2s_{1/2}$ and the $\pi 1d_{3/2}$ orbitals from the shell-model calculations using the SDPF-NR and SDPF-U effective interactions. It is clear that for isotopes from $A = 38 - 46$ and $A =48, 50 - 51$ the dominant component in the configuration is a hole in the $\pi 1d_{3/2}$. In the case of $I^{\pi} = 1/2^{+}$ isotopes, a proton hole is located in the $2s_{1/2}$. Deviation from integer numbers for $^{47-49}$K indicates mixing in the wave function. } \end{figure} The maximum number of particles found in an orbital with total angular momentum $j$ is $2j+1$. Thus, for the $s_{1/2}$ this number is 2, while in case of the $d_{3/2}$ it is 4. The occupation of the $\pi 2s_{1/2}$ remains almost constant around 2 protons from $N=19$ up to $N=27$, with a slight decrease toward $^{46}$K. For these isotopes, the occupation of the $\pi 1d_{3/2}$ stays around 3 protons with a corresponding slight increase toward $A = 46$. This increase (decrease) of occupancy for the $\pi 1d_{3/2}$ ($\pi 2s_{1/2}$) orbital is probably due to the reduction of the energy difference between these two proton orbitals with increasing number of neutrons in the $\nu 1f_{7/2}$. Additionally, a small odd-even staggering in the proton occupation is also observed for these isotopes. This effect could be due to the proton-neutron coupling for the odd-odd isotopes, which results in a higher occupancy of $\pi 1d_{3/2}$ and a lower occupancy for the $\pi 2s_{1/2}$. In this region, there is no discrepancy observed between results from the different interactions. Furthermore, almost degenerate proton orbitals for $N = 28$ yield a hole in the $\pi 2s_{1/2}$ causing the $\pi 1d_{3/2}$ to be nearly completely filled. Surprisingly, for $^{48}$K with an additional unpaired neutron in the $\nu 2p_{3/2}$ orbital, the proton occupation of $\pi 1d_{3/2}$ drops down to about 3.3 protons while the $\pi 2s_{1/2}$ occupation increases accordingly. For the next isotope with two neutrons placed in the $\nu 2p_{3/2}$ ($^{49}$K), the occupation of the proton orbitals is more similar to $^{47}$K, where a hole in the $\pi 2s_{1/2}$ is found. This is also in agreement with the nuclear spins and magnetic moments of these two isotopes. At this point a larger deviation from integer numbers for the proton occupation indicates a larger amount of mixing in the configurations of $^{47\text{-}49}$K. Based on the information obtained from the $g$ factor and magnetic moment for $^{48}$K, a hole in the $\pi d_{3/2}$ was expected to be the dominant component, which is confirmed by these occupancies. Nevertheless, the reason for the big decrease of the $\pi 1d_{3/2}$ occupancy compared to the neighboring two isotopes is still puzzling. Adding one and two more neutrons leads to the "normal" occupation for the neutron-rich K isotopes with the filled $\pi 2s_{1/2}$ and a hole in the $\pi 1d_{3/2}$. \section{Summary} Hyperfine spectra of potassium isotopes between $N = 19$ and $N = 32$ were measured using collinear laser spectroscopy, yielding the nuclear spins and magnetic moments. The experimental results were compared to shell-model calculations using two different effective interactions: SDPF-NR and SDPF-U. Overall good agreement is observed between the measured magnetic moments and theoretical predictions. This allows one to draw conclusions on the composition of the wave function as well as on the proton occupation of the $2s_{1/2}$ and $1d_{3/2}$ orbitals. It was shown that the dominant component of the ground-state wave function for odd-A isotopes up to $^{45}$K arises from a hole in the $\pi 1d_{3/2}$. Additionally, for isotopes with spin $1/2^{+}$ the main component of the wave function is $\pi 2s_{1/2}^{-1}$ with more mixing present in $^{49}$K coming from the almost degenerate $\pi 2s_{1/2}$ and $\pi 1d_{3/2}$ proton orbitals. The nuclear spin of $^{51}$K, which was found to be $3/2$, points to the "normal" ordering of the EPSE, and this is confirmed by the measured magnetic moment that is close to the $\pi 1d_{3/2}$ single particle value. In the case of odd-odd isotopes, the main configuration originates from the coupling of the $\pi 1d_{3/2}^{-1} \otimes \nu (pf)$, and this for all odd-odd isotopes from $N=19$ up to $N = 31$. Only for $^{48}$K, a very fragmented wave function has been observed for the $1^-$ ground state. This level becomes the ground state due to a significant ($>$20\%) contribution from the $\pi 2s_{1/2}^{-1} \otimes \nu (pf)$ configuration. Moreover, a detailed discussion about the evolution of the proton effective single particle energies (ESPEs) was presented. The central term of the monopole interaction was found to have the strongest effect in the changing ESPE beyond $N = 28$. {\it{Ab initio}} calculations of the ESPEs show a considerable decrease (70\%) of the gap between $\pi (1d_{3/2}-2s_{1/2})$ at $N = 28$ presenting a promising starting point for the approach which is currently still under development. The experimental results of the neutron-rich potassium isotopes have a relevant role in the future improvements of the effective shell-model interactions and \textit{ab initio} calculations. Additional experimental data for $^{51}$K and $^{53}$K, in particular the spin of the $^{53}$K ground state and the energy of the $I = 1/2 ^{+}$ states, could provide the final clues about the evolution of the proton $sd$ levels in this region. \begin{acknowledgments} The authors are grateful to F. Nowacki for providing the effective interactions which were used in the shell-model calculations. C.B., T.D. and V.S. would like to thank A. Cipollone and P. Navr\'{a}til for their collaboration on $3N$ forces. This work was supported by the IAP-project P7/12, the FWO-Vlaanderen, GOA 10/010 from KU Leuven, NSF Grant PHY-1068217, BMBF (05 P12 RDCIC), Max-Planck Society, EU FP7 via ENSAR (No. 262010) and STFC Grant No. ST/L005743/1. Gorkov-Green's function calculations were performed using HPC resources from GENCI-CCRT (Grant No. 2014-050707) and the DiRAC Data Analytic system at the University of Cambridge (BIS National E-infrastructure Capital Grant No. ST/K001590/1 and STFC Grants No. ST/H008861/1, ST/H00887X/1, and ST/K00333X/1). We would like to thank to the ISOLDE technical group for their support and assistance. \end{acknowledgments}
\section{Formulation of one-dimensional chain under electric-field} We model a one-dimensional Hubbard model in the Coulomb gauge, as shown in Fig.~\ref{figdia}, with the Hamiltonian \begin{eqnarray} H & = & -\gamma\sum_{\ell\sigma}(d^\dagger_{\ell+1,\sigma}d_{\ell\sigma}+h.c.)+U\sum_\ell \left(n_{\ell\uparrow}-\frac12\right)\left(n_{\ell\downarrow}-\frac12\right) + \sum_{\ell\alpha\sigma}\epsilon_\alpha c^\dagger_{\ell\alpha\sigma}c_{\ell\alpha\sigma}\nonumber \\ & & -\frac{g}{\sqrt{L}}\sum_{\ell\alpha\sigma} (d^\dagger_{\ell\sigma}c_{\ell\alpha\sigma}+h.c.) -\sum_{\ell\sigma}\ell E\left( d^\dagger_{\ell\sigma}d_{\ell\sigma}+\sum_\alpha c^\dagger_{\ell\alpha\sigma}c_{\ell\alpha\sigma}\right), \end{eqnarray} where all orbitals on the $\ell$-th TB site, and their chemical potential, are shifted by $\ell E$. Here we use the unit $e=a=1$. The fermionic chain reservoirs drain the excess energy of the excited electrons on the main tight-binding (TB) lattice. This Hamiltonian in the Coulomb gauge~\cite{diss2} is equivalent to the temporal gauge~\cite{turkowski} and the gauge-covariant form~\cite{aron}. In a long-time limit, we assume we have already reached a well-defined nonequilibrium steady state in the presence of reservoirs. With respect to $d_\ell$, we have two sources for the electronic self-energy, one from the fermion reservoirs and the other from the Coulomb interaction, which we denote as $\Sigma_\Gamma$ and $\Sigma_U$, respectively. Here we make a dynamical mean-field theory (DMFT) assumption that the self-energies are local and identical, except for the energy shift due to the voltage drop along the TB lattice~\cite{satosh}, \begin{equation} G^{r,<}_{\ell\ell}(\omega)=G^{r,<}_{\rm loc}(\omega+\ell E) \mbox{ and } \Sigma^{r,<}_{\ell\ell}(\omega)=\Sigma^{r,<}_{\rm loc}(\omega+\ell E), \end{equation} $\ell=-\infty,\infty$ denotes the lattice site~\cite{diss2}. Here the subscript `loc' refers to the local quantity at the central site $\ell=0$. The Dyson's equation in the steady state for the full retarded Green's function can be expressed in the familiar form as \begin{eqnarray} {\bf G}^r(\omega)^{-1} & = & \left[ \begin{array}{ccccc} \ddots & & & & \\ & \omega-E+{\rm i}\Gamma-\Sigma^r_{U,{\rm loc}}(\omega-E) & \gamma & 0 & \\ & \gamma & \omega+{\rm i}\Gamma-\Sigma^r_{U,{\rm loc}}(\omega) & \gamma & \\ & 0 & \gamma & \omega+E+{\rm i}\Gamma-\Sigma^r_{U,{\rm loc}}(\omega+E) & \\ & & & & \ddots \end{array} \right]\nonumber \\ & = & [\omega+\ell E+{\rm i}\Gamma-\Sigma^r_{U,{\rm loc}}(\omega+\ell E)]\delta_{\ell\ell'}+\gamma\delta_{|\ell-\ell'|,1}. \end{eqnarray} The Weiss-field Green function ${\cal G}$ can be expressed similarly except that the interacting self-energy is omitted at the central site, \begin{equation} [{\cal G}^r(\omega)^{-1}]_{\ell\ell'}=[{\bf G}^r(\omega)^{-1}]_{\ell\ell'}+\Sigma^r_{U,{\rm loc}}(\omega)\delta_{\ell 0}\delta_{\ell' 0}\equiv[{\bf G}^r(\omega)^{-1}+\bm{\Sigma}^r_{U,{\rm loc}}]_{\ell\ell'}, \end{equation} with $\bm{\Sigma}^r_{U,{\rm loc}}={\rm diag}[\cdots,0,0,\Sigma^r_{U,{\rm loc}}(\omega),0,0,\cdots]$. The inversion of the above infinite matrix can be achieved efficiently by a recursive method. We divide the lattice into three parts with the central site $\ell=0$, the left ($\ell=-1,-2,\cdots,-\infty$) and right ($\ell=1,2,\cdots,\infty$) semi-infinite chains. We denote the retarded GF matrix ${\cal F}^r_+$ on the RHS semi-infinite chain as \begin{equation} [{{\cal F}^r_+(\omega)}^{-1}]_{\ell\ell'}= [\omega+\ell E+{\rm i}\Gamma-\Sigma^r_{U,{\rm loc}}(\omega+\ell E)]\delta_{\ell\ell'}+\gamma\delta_{|\ell-\ell'|,1}, \end{equation} with $\ell,\ell'=1,2,\cdots,\infty$. The local GF at the end of the chain ($\ell=1$) [$F^r_+(\omega+E)\equiv{\cal F}^r_+(\omega)_{11}$] can be expressed as a continued fraction \begin{eqnarray} F^r_+(\omega+E) & = & \frac{1}{\omega+E+{\rm i}\Gamma-\Sigma^r_{U,{\rm loc}}(\omega+E)-\frac{\gamma^2}{\omega+2E+{\rm i}\Gamma-\Sigma^r_{U,{\rm loc}}(\omega+2E)-\frac{\gamma^2}{\cdots}}}\nonumber \\ & = & [\omega+E+{\rm i}\Gamma-\Sigma^r_{U,{\rm loc}}(\omega+E) -\gamma^2 F^r_+(\omega+2E)]^{-1},\\ \mbox{ or } F^r_+(\omega+E)^{-1} & = & \omega+E+{\rm i}\Gamma-\Sigma^r_{U,{\rm loc}}(\omega+E) -\gamma^2 F^r_+(\omega+2E), \label{eq:recur1} \end{eqnarray} from the self-similarity of the semi-infinite chain. The recursive relation Eq.~(\ref{eq:recur1}) is solved numerically with iteration number $M$ over 500. Practically, we start from an initial GF $F^r_+(\omega+ME) =[\omega+ME+{\rm i}\Gamma-\Sigma^r_{U,{\rm loc}}(\omega+ME)]^{-1}$ and by Eq.~(\ref{eq:recur1}) we generate $F^r_+(\omega+(M-1)E)$. We repeat the process Eq.~(\ref{eq:recur1}) until we reach $F^r_+(\omega+E)$. The LHS GF, $F^r_-(\omega-E)$, can be similarly obtained through \begin{equation} F^r_-(\omega-E)^{-1} = \omega-E+{\rm i}\Gamma-\Sigma^r_{U,{\rm loc}}(\omega-E) -\gamma^2 F^r_-(\omega-2E). \label{eq:recur2} \end{equation} Once we obtain fully convergent GFs $F^r_{\pm}(\omega\pm E)$, the full local GF for the infinite chain can be constructed as \begin{equation} G^r_{\rm loc}(\omega)^{-1}=\omega+{\rm i}\Gamma-\Sigma^r_{U,{\rm loc}}(\omega) -\gamma^2[F^r_+(\omega+E)+F^r_-(\omega-E)]. \label{eq:gr} \end{equation} The Weiss-field GF ${\cal G}^r(\omega)$, omits the interacting self-energy only on the central site ($\ell=0$) and we have \begin{equation} {\cal G}^r(\omega)^{-1}=\omega+{\rm i}\Gamma-\gamma^2[F^r_+(\omega+E)+F^r_-(\omega-E)] =G^r_{\rm loc}(\omega)^{-1}+\Sigma^r_{U,{\rm loc}}(\omega). \label{eq:wfr} \end{equation} Now, we turn to the Dyson's equation for lesser GFs. When the lattice of $d_\ell$ is connected to the reservoirs and with finite interaction, its steady-state dynamics is governed by the transport equation \begin{equation} G^<_{\ell\ell'}(\omega)= \sum_{p}G^r_{\ell p}(\omega)\Sigma^<_{p,{\rm tot}}(\omega)G^a_{p\ell'}(\omega), \end{equation} with $p=-\infty,\cdots,\infty$ running over all TB sites and $\Sigma^<_{\rm tot}$ being the sum of contributions from the fermion baths and the Hubbard interaction. For the central site $\ell=\ell'=0$, we use a similar trick as above to group $p$ into the central site and left and right chains, \begin{equation} G^<_{\rm loc}(\omega)=G^r_{\rm loc}(\omega)\Sigma^<_{\rm tot,loc }(\omega)G^a_{\rm loc}(\omega)+ \sum_{p<0}G^r_{0p}(\omega)\Sigma^<_{{\rm tot},p}(\omega)G^a_{p0}(\omega) +\sum_{p>0}G^r_{0p}(\omega)\Sigma^<_{{\rm tot},p}(\omega)G^a_{p0}(\omega). \end{equation} For the RHS summation ($p>0$), one can write the Dyson's equation $G^r_{0p}(\omega)=G^r_{\rm loc}(\omega)(-\gamma){\cal F}^r_{+,1p}(\omega)$ and similarly for the advanced FGs, and therefore we have the third term as \begin{equation} \gamma^2|G^r_{\rm loc}(\omega)|^2\sum_{p>0}{\cal F}^r_{+,1p}(\omega)\Sigma^<_{{\rm tot},p}(\omega){\cal F}^a_{+,p1}(\omega). \end{equation} The summed expression is nothing but the local lesser GF $F^<_+(\omega+E)={\cal F}^<_{+,11}(\omega)$ within the LHS semi-infinite chain, and we obtain \begin{equation} G^<_{\rm loc}(\omega)=|G^r_{\rm loc}(\omega)|^2\left\{\Sigma^<_{\rm tot,loc }(\omega)+\gamma^2[F^<_+(\omega+E)+F^<_-(\omega-E)]\right\}. \label{eq:glss} \end{equation} $F^<_\pm(\omega\pm E)$ can be obtained from $\Sigma^<_{\rm loc}(\omega)$ following similar steps. \begin{eqnarray} F^<_+(\omega+E)&=&\sum_{p=1}^\infty {\cal F}^r_{+,1p}(\omega) \Sigma^<_{{\rm tot},p}(\omega){\cal F}^a_{+,p1}(\omega)\nonumber \\ &=&|F^r_+(\omega+E)|^2\Sigma^<_{\rm tot,1}(\omega) +\gamma^2|F^r_+(\omega+E)|^2\sum_{p=2}^\infty \tilde{\cal F}^r_{+,2p}(\omega) \Sigma^<_{{\rm tot},p}(\omega)\tilde{\cal F}^a_{+,p2}(\omega),\nonumber \end{eqnarray} where the tilde denotes that the GFs are on the semi-infinite chains of $p=2,3,\cdots,\infty$. Using the self-similarity of the chains $\ell=1,\cdots,\infty$ and $\ell=2,\cdots,\infty$, we have \begin{equation} F^<_\pm(\omega\pm E)=|F^r_\pm(\omega\pm E)|^2\left[\Sigma^<_{\rm tot,loc} (\omega\pm E)+\gamma^2F^<_\pm(\omega\pm 2E)\right]. \label{eq:flss} \end{equation} The lesser Weiss-field GF can be written as \begin{equation} {\cal G}^<(\omega)=|{\cal G}^r(\omega)|^2\left\{\Sigma^<_{\Gamma,{\rm loc}} (\omega)+\gamma^2[F^<_+(\omega+E)+F^<_-(\omega-E)]\right\}, \end{equation} with the damping part of the self-energy $\Sigma_\Gamma$. Using Eq.~(\ref{eq:glss}), \begin{equation} {\cal G}^<(\omega) = |{\cal G}^r(\omega)|^2\left( \frac{G^<_{\rm loc}(\omega)}{| G^r_{\rm loc}(\omega)|^2}-\Sigma^<_{U,{\rm loc}}(\omega)\right). \label{eq:wflss} \end{equation} In the main paper, the subscript `loc' has been omitted for brevity. Electric current per spin is calculated as \begin{eqnarray} J &=& \frac{i}{2} \gamma\langle d^\dagger_{1\sigma}d_{0\sigma} -d^\dagger_{0\sigma}d_{1\sigma} +d^\dagger_{0\sigma}d_{-1\sigma} -d^\dagger_{-1\sigma}d_{0\sigma}\rangle = \gamma{\rm Re}[G^<_{01}(t=0)-G^<_{0-1}(t=0)]\nonumber \\ & = & \gamma{\rm Re}\int \frac{d\omega}{2\pi} [G^<_{01}(\omega)-G^<_{0-1}(\omega)] \\ & = & -\gamma^2{\rm Re}\int \frac{d\omega}{2\pi} \left\{G^<(\omega)\left[F^a_+(\omega+\Omega)-F^a_-(\omega-\Omega)\right] +G^r(\omega)\left[F^<_+(\omega+\Omega)-F^<_-(\omega-\Omega)\right]\right\}. \nonumber \end{eqnarray} To summarize, given the local self-energies $\Sigma^{r,<}_{\rm loc}(\omega)$, GFs for the semi-infinite chains $F^{r,<}_{\pm}$ are calculated via Eqs.~(\ref{eq:recur1},\ref{eq:recur2},\ref{eq:flss}). The retarded GFs are obtained via Eqs.~(\ref{eq:gr},\ref{eq:wfr}), and finally the lesser GFs follow via Eqs.~(\ref{eq:glss},\ref{eq:wflss}). This procedure, formulated on real-space, corresponds to the $k$-summation of the impurity GF in equailibrium DMFT formalism. \medskip \noindent\textbf{Multi-dimensional lattice under electric-field:} For higher dimensional cubic lattice with the field along an axial direction (${\bf E}=E\hat{\bf x}$), the lattice has translational invariance perpendicular to the field, and the problem is block-diagonalized with the transverse wave-vector ${\bf k}_\perp$. We solve the Dyson's equation as above with the ${\bf k}_\perp$-space (the self-energy $\Sigma^{r,<}_{\rm loc}(\omega)$ does not have ${\bf k}_\perp$ dependence), and then sum over ${\bf k}_\perp$ to get the local GF. For hypercubic TB lattice the dispersion is $\epsilon_{\bf k}=-2\gamma\cos(k_x) +\epsilon({\bf k}_\perp)$. Then adding $\epsilon({\bf k}_\perp)$ to the on-site energy of the 1-$d$ tight-binding chain and carrying out the 1-$d$ Dyson's equation in the previous section, we obtain the GF $G^{r,<}_{{\bf k}_\perp}(\omega)$. By summing over ${\bf k}_\perp$ in the $d-1$ dimensional Brillouin zone, we get the full local GFs \begin{equation} G^{r,<}_{\rm loc}(\omega)=\int_{\rm BZ}\frac{d^{d-1}{\bf k}_\perp}{(2\pi)^{d-1}}G^{r,<}_{{\bf k}_\perp}(\omega) =\int d\epsilon_\perp D_{d-1}(\epsilon_\perp)G^{r,<}(\epsilon_\perp,\omega), \end{equation} with the $d-1$ dimensional DoS $D_{d-1}(\epsilon_\perp)$. The Weiss-field GFs are obtained via Eqs.~(\ref{eq:wfr},\ref{eq:wflss}). \section{Crossover of effective temperature near $U_{\rm cross}\approx 1.32W$} The effective temperature $T_{\rm eff}$ in Fig. 4(c) shows different behavior around $U_{\rm cross}\approx 1.32W$, where $U<U_{\rm cross}$ the increase of $T_{\rm eff}$ with the E-field is rapid in a similar fashion as that well away from the coexisitence region with smaller $U$, whereas for $U>U_{\rm cross}$ the increase of $T_{\rm eff}$ is much slower after the linear response limit. Therefore it leads to a large electric-field to reach the metal-insulator transition near the crossover, resulting in a maximum of the $E_{\rm MIT}(U)$ curve in Fig.~4(a) and (c). Here, we give a sketch of the different scaling behaviors in the two regimes. From the balance of the Joule heating and the dissipation of energy into the fermion baths, we arrive at the rigorous relation Eq.~(39) of Han and Li~\cite{diss2}, \begin{equation} JE=2\Gamma\int\omega A(\omega)[f_{\rm loc}(\omega)-f_{\rm b}(\omega)]d\omega, \label{balance} \end{equation} where $A(\omega)$ is the on-site spectral function of the tight-binding lattice, $f_{\rm b}(\omega)$ the Fermi-Dirac function of the bath. We approximate the local distribution function $f_{\rm loc}(\omega)$ as a Fermi-Dirac function with the effective temperature $T_{\rm eff}$. (i) In the regime of $U<U_{\rm cross}$ with $W^*>T_{\rm eff}\gg T_{\rm b}$, we can apply the Sommerfeld expansion and obtain \begin{equation} JE\approx \frac{\pi^2}{3}\Gamma A(0)(T_{\rm eff}^2-T_{\rm b}^2), \label{phenom} \end{equation} which agrees with the phenomenological energy balance equation of Altshuler et al~\cite{altshuler} with the dissipation given by the fermion baths. Away from the linear response limit, $\tau_U^{-1}\gg\Gamma$ and $J\propto \gamma E/\tau_U^{-1}$ and we have \begin{equation} E^2\propto \Gamma \tau_U^{-1}T_{\rm eff}^2/W^2. \end{equation} From the perturbative self-energy for the scattering rate, as used in the main text, \begin{equation} \tau_{U}^{-1} = -{\rm Im}\Sigma^r_{\rm eq}(\omega=0,T_{\rm eff}) \approx \frac{\pi^3}{2} \, A_0(0)^3 \, U^2 \, T_{\rm eff}^2, \label{eq:scatt} \end{equation} we arrive to the scaling relation, \begin{equation} E^2\propto \Gamma U^2T_{\rm eff}^4/W^5,\mbox{ or } T_{\rm eff}/W\propto (E/U)^{1/2}, \label{relations} \end{equation} with $T_{\rm eff}$ increasing as $\sqrt{E}$ beyond the linear response regime. This scaling agrees well with the numerical calculations. (ii) In the regime of $U>U_{\rm cross}$ with $T_{\rm eff}\lesssim W^*$, the scaling relation is quite different. Although somewhat exaggerated, we assume $T_{\rm eff}\gg W^*$ in the following argument for the sake of simplicity. In such limit, the Sommerfeld expansion is not applicable to both the Eqs.~(\ref{phenom}) and (\ref{eq:scatt}), and the half-QP-bandwidth $W^*/2$ replaces the role of temperature $\pi T_{\rm eff}$, leading to the approximate relations \begin{equation} JE\propto (\Gamma/W){W^*}^2\mbox{ and } \tau_{U}^{-1} \propto U^2 {W^*}^2/W^3, \end{equation} which demonstrates that $T_{\rm eff}$-dependence effectively drops out. Therefore, as shown in the numerical calculations in Fig. 4(c), $T_{\rm eff}$ has much reduced dependency on E-field for $U>U_{\rm cross}$, away from the linear response regime. This slow increase of $T_{\rm eff}$ leads to the enhenced upper switching field $E_{\rm MIT}(U)$ to reach the resistive switching, and the maximum behavior of $E_{\rm MIT}(U)$ results near the crossover value $U\approx U_{\rm cross}$. \begin{figure}[h] \rotatebox{0}{\resizebox{!}{1.7in}{\includegraphics{figS2}}} \caption{Scaling relation of the critical field for the resistive switching vs. the damping parameter $\Gamma$, as the electric-field is increased. The relation relation $E_{\rm MIT}(\Gamma)\propto\sqrt{\Gamma}$ is consistent with the thermal scenario. } \label{figscale} \end{figure} The above scaling relations can be used to derive $E_{\rm MIT}(U)$'s dependence on $\Gamma$. At a given $U$, the nonequilibrium MIT occurs when the effective temperature matches the equilibrium transition temperature $T_{\rm eff}=T_{\rm eq,MIT}(U)$. Then Eq.~(\ref{relations}), for $U<U_{\rm cross}$ leads to \begin{equation} E_{\rm MIT}(U)^2\propto\Gamma U^2T_{\rm eq,MIT}^4/W^5, \mbox{ and } E_{\rm MIT}(U)\propto\sqrt\Gamma. \end{equation} The numerical calculations for $U=1.225$ shown in Fig.~\ref{figscale} confirms the relation.
\section{Introduction} \label{s:Introduction} Recent years have brought remarkable progress in our understanding of the mathematical structure of scattering amplitudes, especially in the planar limit of ${\cal N}=4$ super Yang-Mills theory (SYM). Among the new approaches are the Grassmannian formulations \cite{ArkaniHamed:2009dn,Mason:2009qx,ArkaniHamed:2009vw}, on-shell diagrams \cite{ArkaniHamed:2012nw}, and the geometrization of amplitudes in the ``amplituhedron" \cite{Arkani-Hamed:2013jha,Arkani-Hamed:2013kca,ArkaniHamed:2010gg}. Many of the ideas from planar $\mathcal{N}=4$ SYM carry over to the superconformal 3d ${\cal N}=6$ Chern-Simons matter theory constructed by Aharony, Bergman, Jafferis and Maldacena (ABJM)~\cite{WestCSM}; see also \cite{EastCSM}. In this paper we study the Grassmannian descriptions of amplitudes in both 4d ${\cal N}=4$ SYM and in 3d ABJM theory. The Grassmannian G$(k,n)$ is the set of all $k$-planes in $n$-dimensional space; in the context of scattering amplitudes, $n$ counts the number of external particles while $k$ refers to a classification of amplitudes. The Grassmannian description of amplitudes depends on how the external data --- particle momenta and type --- are encoded. For $n$-particle N$^k$MHV amplitudes in 4d planar $\mathcal{N}=4$ SYM there are three formulations \cite{ArkaniHamed:2009dn,Mason:2009qx}: \begin{itemize} \item The {\bf \em momentum space} formulation uses the spinor helicity representation for the external momenta, e.g.~$p_{a\dot{b}} = \lambda_a \tilde{\lambda}_{\dot{b}}$. The relevant Grassmannian is G$(k+2,n)$. \item The {\bf \em twistor space} formulation encodes the data via a half-Fourier transform of the external momenta, i.e.~a Fourier transform of $\lambda$, but not $\tilde{\lambda}$. This representation makes the superconformal symmetry $SU(2,2|4)$ manifest. As in momentum space, the Grassmannian is G$(k+2,n)$. \item The {\bf \em momentum twistor} formulation is applicable in the planar limit and makes the {\it dual} superconformal symmetry $SU(2,2|4)$ manifest. The relevant Grassmannian is G$(k,n)$. \end{itemize} In Section \ref{s:backgr}, we provide a pedagogical review of these three representations of the 4d external data and present the corresponding Grassmannian integrals explicitly. The three Grassmannian descriptions are directly related. The relation between the 4d SYM twistor space and momentum space integrals was utilized already in the early literature \cite{ArkaniHamed:2009dn} on the subject. The momentum twistor Grassmannian was introduced shortly after in \cite{Mason:2009qx}. Its relation to the two other formulations was given in \cite{ArkaniHamed:2009vw} using a set of intricate integral manipulations. In particular, the argument of \cite{ArkaniHamed:2009vw} uses a gauge fixing that breaks little group scaling and therefore results in very complicated Jacobians that are difficult to write explicitly. In Section \ref{s:TandMT}, we present a new version of the proof, valid for all $n$ and $k$, that manifestly preserves the little group scaling at each step of the calculation and yields all Jacobians as simple explicit expressions. In Section \ref{s:res} we address how the Grassmannian integrals are evaluated as contour integrals and demonstrate this with the explicit computation of the residues in the $n$-point NMHV sector using the momentum twistor Grassmannian integral. The results for the individual residues are known to be the dual superconformal invariant building blocks of the tree amplitudes, i.e.~the ``$R$-invariants" \cite{Drummond:2008vq} or ``5-brackets" \cite{Mason:2009qx}. These invariants obey a set of linear relations, which in the Grassmannian formulation simply follow from global residue theorems. We provide a homological interpretation of the residue theorems for the NMHV residues. We then review the structure of physical versus spurious (unphysical) poles of the residues and how this gives a specification of contours for which the Grassmannian integral exactly produces the tree-level amplitudes. The results are then rephrased in the context of on-shell diagrams and the positroid stratification of the Grassmannian, and we show how the `boundary operation' in the Grassmannian integral is related to both the residue theorems and the pole structure. In Section \ref{sec:3DGrass}, we turn to the Grassmannian descriptions of amplitudes in 3d ABJM theory. We briefly review the momentum space spinor helicity formalism in 3d and the associated Grassmannian integral for ABJM amplitudes, which was introduced previously in \cite{LeeOG,HW,HWX,KimLee}. It encodes the $n{=}(2k{+}4)$-point N$^k$MHV amplitudes of ABJM theory as contour integrals in the {\it orthogonal Grassmannian} OG$(k+2,2k+4)$. The space OG$(k,n)$ is equipped with the metric $g_{ij}=\delta_{ij}$, $i,j=1,2,\dots,n$, and consists of $k$-dimensional null planes in $\mathbb{C}^{n}$; in other words, the $k \times n$ matrices $B \in \text{OG}(k,n)$ satisfy $B g B^T = 0$. A momentum twistor version of the ABJM Grassmannian integral has not previously been constructed. We achieve this goal in Section \ref{sec:3DGrass}, which is an important first step towards developing an amplituhedron for the ABJM theory. To this end, we first introduce another formulation of the 3d spinor helicity formalism that facilitates the definition of 3d momentum twistors. They are simply the 4d momentum twistors $Z_i^A$ ($A=1,2,3,4$) subject to the bi-local $SO(2,3) \sim Sp(4)$-invariant constraint $Z_i^A Z_{i+1}^B \Omega_{AB}=0$ for all $i=1,2,\dots,n$. Next, our streamlined proof of the relation between the 4d Grassmannian integral representations in Section \ref{s:TandMT} allows us to derive the desired momentum twistor version of the Grassmannian integral for ABJM theory. Just like the momentum space Grassmannian integral, the new integral has an orthogonality constraint, but a novel feature is that the $k$-dimensional planes are now null with respect to a metric defined by $Sp(4)$-invariant inner products of the momentum twistors, $Z_i^A Z_j^B \Omega_{AB}$. Orthogonal Grassmannians defined by non-trivial metrics for Grassmannians have been encountered previously in the mathematics literature in the context of electrical networks and related combinatorics \cite{HS,Lam}. However, the dependence of the metric on external data appears to be a new property that would be exciting to explore further. In particular, it plays a crucial role for the boundary properties needed for the physical poles of the 6-point ABJM tree-amplitude, as we discuss in some detail. We end in Section \ref{s:out} with a brief outlook to Grassmannians beyond the NMHV level and open questions. A few technical results are relegated to appendices. \section{${\cal N}=4$ SYM and the Grassmannian} \label{s:backgr} This section is intended as a short review of the Grassmannian formulation of amplitudes in planar ${\cal N}=4$ SYM. Sections \ref{s:extdata1} and \ref{s:extdata2} introduce the basic definitions and concepts needed for (super)amplitudes in ${\cal N}=4$ and present the three different forms of the external data: momentum space, twistor space, and momentum twistor space. See \reef{input} for an overview. Section \ref{s:grassint} presents the Grassmannian integrals. Experts can skip ahead to Section \ref{s:TandMT}. \subsection{External data: momentum space and superamplitudes} \label{s:extdata1} The external data for an amplitude encodes the information about the initial and final state particles in the scattering process; practically we take all states to be outgoing. We are here considering only amplitudes in ${\cal N}=4$ SYM, so each of the $n$ external particles --- labeled $i=1,2,\ldots,n$ --- is specified by a null momentum $p_i$, i.e.~$p_i^2 = 0$, along with a specification of the particle type (e.g.~gluon, gluino, or scalar). The scattering amplitude $A_n$ takes this external data as input and returns a complex number on the support of a delta function that enforces momentum conservation $\delta^4(p_1 + \ldots + p_n)$. In 4d, a null vector $p_i^\mu$ is conveniently written as a $2\times 2$ matrix $p_i^{\dot{a}a}$ with vanishing determinant. Because it has rank 1, the matrix can be expressed as a product of two 2-component vectors:\footnote{The spinor helicity conventions used in this paper are chosen to conform with much of the literature on Grassmannians. They differ from those used in the recent review \cite{Elvang:2013cua}.} $p_i^{\dot{a}a} = \lambda^a \tilde{\lambda}^{\dot{a}}$. Thus, for an $n$-particle amplitude with $n$ external massless particles, the on-shell momenta $p_i$ with $p_i^2=0$ are specified as $\big(\lambda_i,\tilde\lambda_i \big)$. For the purpose of exploring the mathematical properties of amplitudes, it is useful to work with complex-valued momenta. In that case, $\lambda_i$ and $\tilde\lambda_i$ are independent. (Alternatively, we can keep $p_i$ real and work with a metric with signature $(-,-,+,+)$.) The scattering amplitudes are built from Lorentz-invariant contractions of the spinors, such as the angle bracket \begin{equation} \<ij\> := \epsilon_{ab}\lambda_i^a\lambda_j^b\,, \label{defAngle} \end{equation} constructed with the help of the antisymmetric Levi-Civita symbol, here $\epsilon^{12} =-\epsilon^{21} = 1 = -\epsilon_{12} = \epsilon_{21}$, of the $SL(2)$ subgroup of the 4d Lorentz group $SO(3,1)$. Lorentz indices are often suppressed in our presentation. The physical spectrum of ${\cal N}=4$ SYM consists of 16 massless particles: the gluon $g^{\pm}$ with helicity states $h=\pm1$, four gluinos $\Lambda^A$ and $\Lambda_A$ with $h=\pm\tfrac{1}{2}$, and six scalars $S^{AB}$ with $h=0$; $A,B=1,2,3,4$. The helicity $h$ states transform in rank $r=2{-}2h$ fully antisymmetric representations of the global $SU(4)$ $R$-symmetry of ${\cal N}=4$ SYM. It is very convenient to encode the states using anticommuting Grassmann variables $\tilde{\eta}_{iA}$, with fundamental $SU(4)$ index $A=1,2,3,4$ and particle label $i=1,2,\dots,n$; Grassmann monomials are in one-to-one correspondence with the states, e.g.~$\tilde{\eta}_{31} \tilde{\eta}_{33} \tilde{\eta}_{34}$ means that particle 3 is a negative helicity gluino $\Lambda^{134} \sim \Lambda_2$. The $n$-point component amplitudes $A_n$ combine into {\bf \em superamplitudes} $\mathcal{A}_n^\text{N$^k$MHV}$ (or more generally $\mathcal{A}_n$), which are polynomials of degree $4(k+2)$ in the Grassmann variables. $R$-symmetry requires $\mathcal{A}_n$ to be an $SU(4)$ singlet, hence the Grassmann degree of each term must be a multiple of 4. The label N$^k$MHV stands for (Next-to)$^k$ Maximally Helicity Violating --- this sector of amplitudes consists of all gluon amplitudes with $k+2$ negative helicity gluons and $n-k-2$ positive helicity gluons, as well as all amplitudes related to those via supersymmetry. The sector with $k=0$ is simply called MHV. The coefficient of a given Grassmann monomial in $\mathcal{A}_n$ is a component amplitude whose external states are those dictated by the Grassmann variables; for example, the coefficient of the monomial $(\tilde{\eta}_{11} \tilde{\eta}_{12} \tilde{\eta}_{13} \tilde{\eta}_{14})( \tilde{\eta}_{23} \tilde{\eta}_{24})(\tilde{\eta}_{41} \tilde{\eta}_{42})$ is the component amplitude $A_n\big[g^-(p_1) S^{34}(p_2) g^+(p_3) S^{12}(p_4) g^+(p_6) \ldots g^+(p_n)\big]$. The $SU(N)$ gauge group of ${\cal N}=4$ SYM dresses the amplitudes with a color-structure that factorizes from the kinematic information. Amplitudes in the {\em planar} theory have a single trace of $SU(N)$ generators,\footnote{For further details, see Section 2.5 of the review \cite{Elvang:2013cua}.} and as a result the planar $n$-particle superamplitudes are invariant under {\em cyclic permutations} of the external labels, i.e.~under $i \rightarrow i+1$ mod $n$. The MHV sector is the simplest. The tree-level MHV superamplitude is given by the supersymmetrization of the Parke-Taylor gluon amplitude \cite{Parke:1986gb,Nair:1988bq}: \begin{equation} \mathcal{A}_n^\text{MHV} = \frac{\delta^4\big( \sum_{i=1}^n\lambda_i \tilde{\lambda}_i \big) \, \delta^{(8)}\big( \sum_{i=1}^n \lambda_i \tilde{\eta}_i \big)} {\<12\> \<23\> \cdots \<n1\>}\,. \label{AMHV} \end{equation} The four bosonic delta functions in \reef{AMHV} encode momentum conservation via $ p_i=\lambda_i \tilde{\lambda}_i$, while the Grassmann delta function,\footnote{\label{footie:delta}The bosonic delta functions are defined as standard in distribution theory \cite{MR0435831}, i.e.~they have the property that $\int_{\mathbb{R}^m} d^m x\,\delta^m(x-x_0) \,f(x) = f(x_0)$ for any suitable test function $f$. Similarly, we use $\int d^{m}x\, d^{n}y \, \delta^{n}\big( g(x,y) \big) f(x,y) = \int d^m x\, \sum_{g(x,y)=0} f(x,y)/\det(dg/dy)$. The Grassmann delta functions are defined by the same property, $\int d\eta \, \delta^{(1)}(\eta-\eta_0) \, f(\eta) = f(\eta_0)$, but using the Berezin integral $\int d\eta \, \eta = 1$ and $\int d\eta \, 1 = 0$. Thus, the Grassmann delta function is simply $\delta^{(1)}(\eta) = \eta$ and $\delta^{(2)}\big(\sum_i\lambda_i \eta_i\big) = \tfrac{1}{2}\sum_{i,j} \<i j\> \eta_{i} \eta_{j}$. The superscript on the Grassmann delta-function indicates its polynomial Grassmann degree. } defined as \begin{equation} \delta^{(8)}\Big( \sum_{i=1}^n \lambda_i \tilde{\eta}_i \Big) := \frac{1}{2^4} \prod_{A=1}^4 \sum_{i,j=1}^{n} \<ij\>\, \tilde{\eta}_{iA} \tilde{\eta}_{jA} \,, \end{equation} ensures conservation of ${\cal N}=4$ supermomentum, $q_{iA} := \lambda_i \tilde{\eta}_{iA}$. The superamplitude \reef{AMHV} clearly has cyclic symmetry. To summarize, for $n$-particle superamplitudes in ${\cal N}=4$ SYM, the external data is specified in terms of the set $\big(\lambda_i,\tilde\lambda_i \,|\, \tilde{\eta}_{iA}\big)$ for $i=1,2,\dots,n$. We call this the {\bf\em momentum space} representation of the external data (or sometimes `on-shell superspace'). The momentum space external data has a redundancy known as {\bf \em little group scaling}: \begin{align} \lambda_i\rightarrow t_i\lambda_i\,,~~~~~ \tilde\lambda_i \rightarrow t_i^{-1}\tilde\lambda_i\,,~~~~~ \tilde{\eta}_i \rightarrow t_i^{-1}\tilde{\eta}_i\,, \label{litgrp} \end{align} for each $i=1,2,\dots,n$. A component amplitude scales homogeneously under little group scaling with weight $t_i^{-2h_i}$, where $h_i$ is the helicity of the $i^\text{th}$ particle. The scaling of the Grassmann variables ensures uniform weight for all external states in a superamplitude: \begin{equation} \label{litgrpAmp} \mathcal{A}_n \rightarrow t_i^{-2} \mathcal{A}_n \,. \end{equation} for each $i=1,2,\dots,n$. Little group scaling plays a key role in several explorations of scattering amplitudes, including the work we present in this paper. \subsection{External data: twistor space and momentum twistor space} \label{s:extdata2} In addition to the momentum space representation, we will be using two other formulations for the 4d external data, namely twistor space and momentum twistor space. We describe each in turn. {\bf \em Twistor space} is obtained from momentum space via a Fourier transform of $\lambda_j$, formally via \begin{equation} \int d^2 {\lambda}_j \, \exp(-i \tilde{\mu}^{{a}}_j{\lambda}_{j{{a}}})\, \bullet \,, \label{FTdef} \end{equation} for each $j=1,2,\ldots,n$. The bullet indicates the expression that is Fourier transformed. (We are ignoring factors of $2\pi$ in all Fourier transforms here and henceforth as these only amount to overall normalizations.) The external data is then encoded in the 4-component twistor $W_i = (\tilde{\mu}_i,\tilde{\lambda}_i)$ and its companion, the supertwistor $\mathcal{W}_i= (\tilde{\mu}_i,\tilde{\lambda}_i\, |\, \tilde{\eta}_i)$. Under little group scaling \reef{litgrp}, we have $\tilde{\mu}_i \rightarrow t_i^{-1} \tilde{\mu}_i$, so the (super)twistor scales uniformly, e.g.~$\mathcal{W}_i \rightarrow t_i^{-1} \mathcal{W}_i$. The measure in the integral \reef{FTdef} scales as $t_i^{2}$, so this exactly compensates the little group scaling of the superamplitude \reef{litgrpAmp}. Thus, after the half-Fourier transformation for all $j=1,2,\ldots,n$, the superamplitude is invariant under little group scaling. In other words, the superamplitude in twistor space is defined projectively, and the twistors $W_i$ and supertwistors $\mathcal{W}_i$ are homogeneous coordinates of projective space, $\mathbb{CP}^3$ and $\mathbb{CP}^{3|4}$, respectively. The third description of the external data uses the 4-component {\bf \em momentum twistors} $Z_i = (\lambda_i, \mu_i)$ \cite{Hodges:2009hk} and their momentum supertwistor extensions $\mathcal{Z}_i = (\lambda_i, \mu_i \,|\,\eta_i)$. The 2-component spinors $\mu_i$ are defined via incidence relations\footnote{For a more comprehensive review of dual space and momentum twistors, see Section 5.4 of \cite{Elvang:2013cua}.} \begin{equation} \mu_i := \lambda_i y_i = \lambda_i y_{i+1}\,, \label{incidence4d} \end{equation} where the dual space coordinates $y_i$ are defined in terms of the momenta as \begin{equation} p_i = y_i - y_{i+1}\,. \label{pandy} \end{equation} The second relation in \reef{incidence4d} follows from the Weyl equation, $p_i \lambda_i = 0$. The definition \reef{pandy} makes momentum conservation automatic via the identification $y_{n+1}=y_1$. The on-shell condition $p_i^2=0 $ requires `adjacent' points $y_i$ and $y_{i+1}$ to be null separated. Dual conformal symmetry acts on the dual space variables $y_i$ in the familiar way, e.g.~under dual inversion we have $y_i \rightarrow y_i/y_i^2$. The geometric interpretation of the incidence relations \reef{incidence4d} is that a point $Z_i = (\lambda_i, \mu_i)$ in momentum twistor space corresponds to a null line defined by the points $y_i$ and $y_{i+1}$ in dual space. Similarly, the line defined by $Z_{i-1}$ and $Z_i$ in momentum twistor space maps to a point in dual space via \begin{equation} y_i = \frac{\lambda_i \mu_{i-1}-\lambda_{i-1} \mu_{i} }{\<i-1,i\>} \,. \label{y-lambda-mu} \end{equation} This follows from \reef{incidence4d}. In our applications, we need to be able to map directly from momentum space variables $(\lambda_i, \tilde\lambda_i \,|\, \tilde{\eta}_i)$ to momentum twistor variables $\mathcal{Z}_i = (\lambda_i, \mu_i \,|\,\eta_i)$. This is done via the relations \begin{equation} \begin{split} \tilde{\lambda}_{i} &=~ \frac{\two{i+1,i}\mu_{i-1} + \two{i,i-1}\mu_{i+1} + \two{i-1,i+1}\mu_{i} } {\two{i-1,i}\two{i,i+1}}\,, \\[2mm] \tilde{\eta}_{iA} &=~ \frac{\two{i+1,i}\eta_{i-1,A} + \two{i,i-1}\eta_{i+1,A} + \two{i-1,i+1}\eta_{iA} } {\two{i-1,i}\two{i,i+1}}\,. \end{split} \label{inc} \end{equation} It follows from \reef{inc} that both $\mu_{i}$ and $\eta_i$ scale linearly with $t_i$ under little group transformations, so the momentum (super)twistors scale uniformly, e.g.~$\mathcal{Z}_i \rightarrow t_i \mathcal{Z}_i$. Therefore, the $\mathcal{Z}_i$ naturally live in projective space, $\mathbb{CP}^3$ and $\mathbb{CP}^{3|4}$. With the external data given in momentum twistor space, the superamplitude still scales uniformly as in \reef{litgrpAmp}. However, as we shall see, one can split off the MHV superamplitude \reef{AMHV} as an overall factor; it takes care of the scaling properties and leaves behind an object that is invariant under little group scaling and therefore projectively well-defined. The relations between the three different forms of the external data can be summarized compactly as follows: \begin{equation} \hspace{-2mm} \boxed{ \begin{array}{ccccccc} \text{twistor space} &\multicolumn{3}{c}{\hspace{6mm}\text{momentum space}} & \text{momentum twistor space}\\[3mm] \mathcal{W}_i = (\tilde{\mu}_i,\tilde{\lambda}_i \, | \, \tilde{\eta}_i) &\!\!\longleftrightarrow\!\!& (\lambda_i,\tilde{\lambda}_i \, | \, \tilde{\eta}_i) ~~ &\!\!\longleftrightarrow\!\!& \mathcal{Z}_i = (\lambda_i, \mu_i \, | \, \eta_i)\\[-2mm] & \!\!\text{\tiny Fourier transform}\!\! && \!\!\!\!\!\!\!\!\!\text{\tiny incidence relations}\!\!\!\!\!\!\!\!\! \\[-3mm] & \text{\tiny eq \reef{FTdef}} && \text{\tiny eq \reef{inc}} \end{array}} \label{input} \end{equation} Superamplitudes in ${\cal N}=4$ SYM enjoy superconformal symmetry $SU(2,2|4)$; the action of this symmetry is linearized in (super)twistor variables. In the planar limit, the superamplitudes (at tree-level or more generally the loop-{\em integrands}) also have dual superconformal symmetry $SU(2,2|4)$ whose action is linearized in the momentum twistor description. The generators of the `ordinary' and dual superconformal symmetries can be arranged to generate an infinite-dimensional algebra called the $SU(2,2|4)$ Yangian. Further details of the representation of the amplitudes and their symmetries can be found in \cite{Elvang:2013cua}. \subsection{Grassmannian integrals} \label{s:grassint} The complex Grassmannian G$(k,n)$ is the space of $k$-planes in $\mathbb{C}^n$. A $k$-plane can be described as a collection of $k$ $n$-component vectors. Since any $GL(k)$ rotation of the vectors yield the same $k$-plane, the Grassmannian G$(k,n)$ can be given equivalently in terms of $k \times n$ matrices modulo $GL(k)$. The dimension of G$(k,n)$ is therefore $kn -k^2 = k(n-k)$. Here, we will list and briefly describe the three Grassmannian integrals relevant for amplitudes in $\mathcal{N}=4$ SYM; the actual connection to the amplitudes is made in Section \ref{s:res}. \subsubsection{Grassmannian with twistor space data $\mathcal{W}$} In terms of twistor variables, $\mathcal{W}$, the relevant Grassmannian integral was first presented in \cite{ArkaniHamed:2009dn}. For the N$^k$MHV sector of $n$-point superamplitudes, the associated Grassmannian is G$(k+2,n)$ and in this space we study the integral \begin{equation} \tilde{\mathcal{L}}_{n;k}(\mathcal{W}) = \int \frac{d^{\tilde{k}\times n} B}{GL(\tilde{k})}\, \frac{\delta^{4\tilde{k}|4\tilde{k}}\big(B\cdot\mathcal{W}\big)}{m_1 m_2 \dots m_n}\,. \label{GrassW} \end{equation} Here $\tilde{k} = k+2$ and the $m_i$'s are the $\tilde{k} \times \tilde{k}$ consecutive minors of the matrix $B$, i.e.~$m_1 = (1\,2\ldots \tilde{k})_B$, $m_2 = (2\, 3 \ldots \tilde{k}+1)_B$, $\dots$, $m_n = (n\,1\ldots \tilde{k}-1)_B$. The integral \reef{GrassW} should be understood as a contour integral; this will be discussed in Section \ref{s:contours} and more concretely in Section \ref{s:NMHVeval}.\footnote{The integral \reef{GrassW} exhibits two conventions typical of this field. First, we write an integral over a parameter space with $r$ complex parameters to mean that the integral will be taken over a real $r$-dimensional contour to be specified later. Second, let $X$ be a parameter space on which some connected group $G$ acts, let $d X$ be a $G$-invariant volume form on $X$ and choose a left invariant volume form $\mu_G$ on $G$. We write $d X/G$ for the volume form on $X/G$ so that, if we locally identify a patch on $X$ with a product of a patch on $X/G$ and a patch on $G$, then $dX = (dX/G) \times \mu_G$. We do not actually specify the measure $\mu_G$, since it only adds a global constant factor.} The external data enters the integral \reef{GrassW} only via the argument of the delta-functions $B \cdot \mathcal{W} = \sum_{i=1}^n B_{\alpha i}\mathcal{W}_i$ with $\alpha = 1,2, \dots, \tilde{k}$. Specifically, we have \begin{equation} \delta^{4\tilde{k}|4\tilde{k}}\big(B\cdot\mathcal{W}\big) = \prod_{\alpha=1}^{\tilde{k}} \delta^{4}\Big( \sum_i B_{\alpha i} W_i\Big) \, \delta^{(4)}\Big( \sum_i B_{\alpha i} \tilde{\eta}_i\Big) \, \end{equation} with the sum over $i=1,2,\ldots, n$. Note two simple properties of \reef{GrassW}: \begin{itemize} \item Little group scaling, $\mathcal{W}_i \rightarrow t_i^{-1} \mathcal{W}_i$, can be absorbed via a scaling of the $i^\text{th}$ column of $B$: $B_{\alpha i} \rightarrow t_i B_{\alpha i}$ for all $\alpha =1,\ldots,\tilde{k}$. The $i^\text{th}$ column is included in exactly $\tilde{k}$ minors, so the scaling of the product of minors is $t_i^{\tilde{k}}$ and this precisely cancels the scaling of the measure $d^{\tilde{k}\times n}B$. Thus, $\tilde{\mathcal{L}}_{n;\tilde{k}}$ is invariant under little group scaling; it is projectively defined, just as are the superamplitudes in twistor space. \item $\tilde{\mathcal{L}}_{n;\tilde{k}}$ produces objects of Grassmann degree $4 \tilde{k}=4(k+2)$ which is the same as for superamplitudes in the N$^{k}$MHV sector. \end{itemize} \subsubsection{Grassmannian with momentum space data $(\lambda,\tilde{\lambda}\,|\,\tilde{\eta})$} \label{s:grassmMom} In momentum space, the Grassmannian for $n$-point N$^k$MHV amplitudes is also G$(k+2,n)$. The integral can be written \begin{equation} {\mathcal{L}}_{n;k}\big(\lambda,\tilde{\lambda},\tilde{\eta}\big) = \int \frac{d^{\tilde{k}\times n}B}{{GL}(\tilde{k})}\, \frac{ \delta^{2\tilde{k}}\big(B_{\alpha i}\, \tilde{\lambda}_i\big)\, \delta^{2(n-\tilde{k})}\big({B}^\perp_{\alpha i} \lambda_i\big)\, \delta^{(4\tilde{k})}\big(B_{\alpha i}\,\tilde{\eta}_{iA}\big) }{m_1 m_2 \cdots m_n} \,,\\ \label{GrassM} \end{equation} where $\tilde{k}=k+2$ and $B^\perp$ is the $(n-\tilde{k})\times n$ matrix parameterizing the $(n-\tilde{k})$-plane orthogonal to the $\tilde{k}$-plane defined by $B$; i.e.~$B(B^\perp)^T = 0$.\footnote{$B^\perp$ is defined only up to a $GL(n-\tilde{k})$ redundancy, but after fixing the $GL(\tilde{k})$ of $B$ we can choose a canonical $B^\perp$ to avoid ambiguities.} The momentum space Grassmannian integral \reef{GrassM} has $2n$ bosonic delta-functions whereas the twistor space version \reef{GrassW} has $4(k+2)$; the difference arises from the Fourier transformations that relate \reef{GrassW} and \reef{GrassM}, as we review in detail in Section \ref{s:TandMT}. The first $2\tilde{k}$ delta functions in \reef{GrassM} require that the 2-plane defined by the $n$ $\tilde{\lambda}_i$'s must lie in the orthogonal complement to the $\tilde{k}$-plane defined by $B$. The remaining $2(n-\tilde{k})$ delta functions require the $\lambda$ 2-plane to be in the orthogonal complement of $B^\perp$; i.e.~the $\lambda$-plane must be contained in $B$. Hence, the bosonic delta functions require the 2-planes defined by $\lambda$ and $\tilde{\lambda}$ to be orthogonal: $\sum_i \lambda_i \tilde{\lambda}_i = 0$. This is just momentum conservation. We conclude that 4 of the $2n$ bosonic delta-functions in \reef{GrassM} simply enforce a condition on the external data, thus leaving constraints only on $2n-4$ of the integration variables $B$. \subsubsection{Grassmannian with momentum twistor space data $\mathcal{Z}$} The G$(k,n)$ Grassmannian integral with external data given in momentum twistor space was introduced in \cite{Mason:2009qx}. For the N$^k$MHV sector with $n$ external particles it is \begin{equation} \mathcal{L}_{n;k}(\mathcal{Z}) = \mathcal{A}_n^{\text{MHV}} \int \frac{d^{k\times n} C}{GL(k)}\,\frac{\delta^{4k|4k}\big(C\cdot\mathcal{Z}\big)}{M_1 M_2 \cdots M_n}\,. \label{GrassZ} \end{equation} The $k \times k$ minors of the matrix $C$ are $M_1 = (1\,2\ldots k)_C$, etc., and the overall factor is the MHV superamplitude \reef{AMHV}. As above we note that \begin{itemize} \item The integral on the RHS of \reef{GrassZ} is invariant under little group scaling $\mathcal{Z}_i \rightarrow t_i \mathcal{Z}_i$ after a compensating scaling by $t_i^{-1}$ of the $i^\text{th}$ column of $C$. However, the MHV factor scales as $t_i^{-2}$. Thus $\mathcal{L}_{n;k}(\mathcal{Z}) \rightarrow t_i^{-2}\mathcal{L}_{n;k}(\mathcal{Z})$; this is precisely the scaling \reef{litgrpAmp} needed for superamplitudes in momentum twistor space. \item ${\mathcal{L}}_{n;{k}}$ produces objects of Grassmann degree $4 k + 8$, with the ``$+8$" arising from the MHV factor. This is the correct count for superamplitudes in the N$^{k}$MHV sector. \end{itemize} \subsubsection{Contours} \label{s:contours} Beyond the comments about little group scaling and Grassmann degrees, we have not yet established the connection between the Grassmannian integrals and superamplitudes. The first step is to define what is actually meant by the integrals. The idea is the same in all three cases, so we focus on the momentum twistor integral \reef{GrassZ}. Fixing the $GL(k)$ invariance of ${\mathcal{L}}_{n;{k}}$ in \reef{GrassZ} leaves an integral over $k(n-k)$ variables. Of these, the bosonic delta functions localize $4k$. Thus, we are left with $k(n-k-4)$ variables to be integrated. The prescription is to interpret the integrals as $k(n-k-4)$-dimensional contour integrals. We can consider contours that select $k(n-k-4)$ simultaneous zeros of the minors. For each such contour $\gamma$, the integral \reef{GrassZ} computes a $k(n-k-4)$-dimensional residue ${\mathcal{L}}_{n;{k}}^{(\gamma)}$. The sum of certain sets of such residues turns out to be exactly the N$^k$MHV tree superamplitude in momentum twistor space: denoting the corresponding contour $\Gamma_\text{tree}$, we therefore have ${\mathcal{L}}_{n;{k}}^{(\Gamma_\text{tree})}(\mathcal{Z}) = \mathcal{A}_{n,\text{tree}}^\text{N$^k$MHV}(\mathcal{Z})$. We demonstrate the explicit calculation of the individual NMHV residues in Section \ref{s:res} and discuss the associated global residue theorems in Section \ref{s:resrels}. The NMHV `tree-contour' $\Gamma_\text{tree}$ is described in Section \ref{s:resapp}. The Grassmannian integrals \reef{GrassW}, \reef{GrassM}, and \reef{GrassZ} are directly related. This was argued in \cite{ArkaniHamed:2009vw} and we now provide a streamlined proof. \section{Relating the three Grassmannian formulations} \label{s:TandMT} The twistor space and momentum space Grassmannian integrals \reef{GrassW} and \reef{GrassM} are easily related via the half-Fourier transform \reef{FTdef}; for completeness we review this below. The derivation of the momentum twistor Grassmannian integral \reef{GrassZ} from either of the other two integrals requires more effort since one needs to reduce the Grassmannian G$(k+2,n)$ to G$(k,n)$. As noted in the Introduction, this was first done in \cite{ArkaniHamed:2009vw}. We present here a streamlined and more explicit version of the proof; this will be useful for deriving the equivalent momentum twistor Grassmannian integral for ABJM theory in Section \ref{sec:3DGrass}. \subsection{From twistor space to momentum space} The Grassmannian integral in twistor space $\widetilde{\mathcal{L}}_{n;k}\big(\mathcal{W}\big)$ is converted to momentum space via the inverse of the Fourier transform \reef{FTdef} that relates momentum space and twistor space. Thus, the momentum space Grassmannian integral is given as \begin{equation} {\mathcal{L}}_{n;k}\big(\lambda,\tilde{\lambda},\tilde{\eta}\big) = \bigg( \prod_{i=1}^n \int d^2 \tilde\mu_i\, \,e^{i {\lambda}_i.\tilde{\mu}_i}\bigg) \widetilde{\mathcal{L}}_{n;k}\big(\mathcal{W}\big)\, . \label{FTofLnk} \end{equation} Since $\widetilde{\mathcal{L}}_{n;k}\big(\mathcal{W}\big)$ is invariant under little group scaling, the expression ${\mathcal{L}}_{n;k}$ scales as $t_i^{-2}$ thanks to the scaling of the measure of the Fourier transform. The only $\tilde\mu$-dependent part of $\widetilde{\mathcal{L}}_{n;k}\big(\mathcal{W}\big)$ is $\delta^{2\tilde{k}}\big(B\cdot\tilde\mu\big)$, as can be seen from \reef{GrassW}. This $\delta$-function enforces that $B$ must be orthogonal to the 2-plane defined by $\tilde\mu_a$ (viewed as two $n$-component vectors). It is convenient to introduce the $B^\perp$ as the $(n-\tilde{k})\times n$ matrix parameterizing the $(n-\tilde{k})$-plane orthogonal to the $\tilde{k}$-plane defined by $B$; i.e.~it satisfies $B(B^\perp)^T = 0$. The constraints of $\delta^{2\tilde{k}}\big(B\cdot\tilde\mu\big)$ can then be reformulated as $\tilde\mu \subset B^\perp$. In other words, $\tilde\mu_a$ is some linear combination of the rows of $B^\perp$: \begin{equation} \delta^{2k}\big(B_{\alpha i}\tilde\mu_i\big) = \int d^{2(n-\tilde{k})}\sigma_{\bar\alpha}\, \delta^{2n}\big(\tilde\mu_i - \sigma_{\bar\alpha} B^\perp_{\bar\alpha i}\big)\,, \label{sigmaInt} \end{equation} where $\bar\alpha = 1,\ldots,n-\tilde{k}$. We can now easily perform the inverse-Fourier transform back to momentum space. The delta functions \reef{sigmaInt} localize the Fourier integral \reef{FTofLnk} to give $e^{i \sigma \cdot B^\perp\cdot \lambda}$, so that integration of the $\sigma$'s then yields $2(n-\tilde{k})$ new delta functions $\delta^{2(n-\tilde{k})} \big( B^\perp \cdot \lambda\big)$. The result is the momentum space Grassmannian integral \reef{GrassM}. \subsection{Derivation of the momentum twistor Grassmannian} Having derived the momentum space integral \reef{GrassM} from the twistor space one \reef{GrassW}, we now continue to momentum twistor space. The key step is the reduction of the integral from G$(k+2,n)$ to G$(k,n)$. The bosonic delta functions $\delta^{2(n-\tilde{k})} \big( B^\perp \cdot \lambda\big)$ in \reef{GrassM} require that the $\lambda$ 2-plane lies in the orthogonal complement of $B^\perp$, so \begin{equation} \prod_{\beta=1}^{\tilde{k}} \int d^2\rho_\beta \,\delta^{2n} \big( \lambda_j - \rho_\alpha B_{\alpha j}\big) \,, \label{2ndeltas} \end{equation} where $\rho^a_{\alpha}$ is a $2 \times \tilde{k}$ array of dummy integration variables. As an aside, let us note that we could easily have found \reef{2ndeltas} directly from the inverse Fourier integral of the twistor space integral \reef{FTofLnk} by writing $\delta^{2\tilde{k}}\big(B\cdot \tilde{\mu}\big)$ as $\int d^2\rho_\alpha\,e^{-i\rho_\alpha B_{\alpha j} \tilde{\mu}_j}$ and then carrying out the $2n$ Fourier integrals in \reef{FTofLnk} to find \reef{2ndeltas}. The $GL(\tilde{k})=GL(k+2)$ redundancy of the $B$'s is transferred to the $\rho$'s. So we can go ahead and fix part of $GL(k+2)$ by choosing \begin{equation} \rho = \left( \begin{array}{ccccccc} 0 &\cdots &0 & 1 & 0\\ 0 &\cdots &0 & 0 &1 \end{array} \right)\,. \end{equation} The $2n$ delta functions \reef{2ndeltas} then fix the last two rows of $B$ to be the $\lambda$'s: \begin{equation} B=\left( \begin{array}{cccc} B_{11} & B_{12} & \cdots & B_{1n} \\ \vdots & \vdots & \ddots & \vdots \\ B_{k1} & B_{k2} & \cdots & B_{kn} \\ \lambda_{1}^1 & \lambda_{2}^1 & \cdots & \lambda_{n}^1 \\ \lambda_{1}^2 & \lambda_{2}^2 & \cdots & \lambda_{n}^2 \end{array} \right)\,. \label{Cfixed} \end{equation} Thus, after evaluating the $\rho$ integrals, we find \begin{equation} \label{momgrass} {\mathcal{L}}_{n;k}\big(\lambda,\tilde{\lambda},\tilde{\eta}\big) = \delta^4\big(\lambda_i \tilde\lambda_i\big) \delta^{(8)}\big(\lambda_i \tilde{\eta}_i\big) \times \int \frac {d^{k\times n} B_{\hat{\alpha} i}} {GL(k)\ltimes T_k} \, \frac {\delta^{2k}\big(B_{\hat\alpha i}\tilde\lambda_i\big) \, \delta^{(4k)}\big(B_{\hat\alpha i}\tilde{\eta}_i\big)} {m_1 m_2 \cdots m_n}\,, \end{equation} with $\hat{\alpha}=1,2,\ldots,k$. The gauge choice \reef{Cfixed} preserves little group scaling. Note that all the delta functions in \reef{momgrass} are little group invariant using $B_{\hat\alpha i} \rightarrow t_i B_{\hat\alpha i}$. Again, the $n$ minors scale as $t_i^{\tilde{k}}=t_i^{k+2}$, but now the measure only contributes $t_i^{k}$. So overall, the expression \reef{momgrass} for ${\mathcal{L}}_{n;k}$ scales as $t_i^{-2}$, as anticipated. In \reef{momgrass}, $T_k$ indicates the translational redundancy in the $B_{\hat\alpha i}$-variables. The translational symmetry acts as \begin{equation} B_{\hat\alpha i} \rightarrow B_{\hat\alpha i} + r_{1\hat\alpha} \lambda_i^1 + r_{2\hat\alpha} \lambda_i^2 ~~~~~ \text{for all $i$ simultaneously,} \label{translation} \end{equation} where $r_{1\hat\alpha}$ and $r_{2\hat\alpha}$ are any numbers. This is a mixing of the last two rows in the $B$-matrix \reef{Cfixed} with the other rows, and this leaves the minors $m_i$ unchanged. It is also clear that on the support of the two delta functions $\delta^4\big(\lambda_i \tilde\lambda_i\big) \delta^{(8)}\big(\lambda_i \tilde{\eta}_i\big) $ (that encode momentum and supermomentum conservation), the delta-functions in the integral \reef{momgrass} are invariant under such a shift. So far, what we have done parallels the work \cite{ArkaniHamed:2009vw}. At this stage, the authors of \cite{ArkaniHamed:2009vw} fix the translation invariance $T_k$ via $2k$ delta functions $\delta(B_{\hat\alpha i } \lambda_i)$. This breaks the little group scaling and the associated Jacobian is therefore unpleasant. We proceed here in a way that preserves little group scaling at every step and gives very simple Jacobians that can be presented explicitly. We change variables in the external data to go from momentum space to momentum twistor space. The momentum supertwistors $\mathcal{Z}_i = (\lambda_i, \mu_i \,|\, \eta_i )$ are related to the momentum space variables via the relations \reef{inc}. Using these relations, we directly find for each $\hat\alpha=1,2,\dots,k$: \begin{equation} \sum_{i=1}^n B_{\hat\alpha i}\tilde{\lambda}_i = - \sum_{i=1}^n C_{\hat\alpha i}\mu_i\,, ~~~~~~~~~ \sum_{i=1}^n B_{\hat\alpha i}\tilde{\eta}_i = - \sum_{i=1}^n C_{\hat\alpha i}\eta_i\,, \end{equation} where the reorganization on the RHS directly gives \begin{equation} \label{DfromC} C_{\hat\alpha i} = \frac{\two{i,i+1}B_{\hat\alpha, i-1} + \two{i-1,i}B_{\hat\alpha, i+1}+ \two{i+1,i-1}B_{\hat\alpha i} } {\two{i-1,i}\two{i,i+1}}\,. \end{equation} A sign was absorbed which flipped the angle brackets relative to \eqref{inc}. The expression \reef{DfromC} implies that $C_{\hat\alpha i} \rightarrow t_i^{-1} C_{\hat\alpha i}$ under little group scaling. We can rewrite the $(k+2) \times (k+2)$ minors $m_i$ of the $B$-matrix in terms of the $k\times k$ minors of the $C$-matrix as \cite{ArkaniHamed:2009vw} \begin{equation} m_1 = (B_1 \dots B_{k+2}) = -\<12\> \cdots \<k+1,k+2\> (C_2\dots C_{k+1} )\,~~~\text{etc.} \label{minors} \end{equation} Defining the $k \times k$ minors of the $k\times n$ $C$-matrix to be $M_1 := (C_1\dots C_{k})$ etc, we thus have \begin{equation} m_1 m_2 \cdots m_n = (-1)^n \big( \two{12}\two{23} \cdots \two{n1} \big)^{k+1} M_1 M_2 \cdots M_n \,. \label{minorleague} \end{equation} (We drop the signs $(-1)^n$ just as we drop $2\pi$'s in the Fourier transforms.) Thus, we now have \begin{equation} \label{momgrass2} {\mathcal{L}}_{n;k}\big(\lambda,\tilde{\lambda},\tilde{\eta}\big) = \frac{\delta^{4}\big(\lambda_i \tilde\lambda_i\big) \,\delta^{(8)}\big(\lambda_i \tilde{\eta}_i\big) } {\big( \two{12}\two{23} \cdots \two{1n} \big)^{k+1}} \int \frac{d^{k\times n} B_{\hat\alpha i}} {GL(k)\ltimes T_k} \, \frac{\delta^{2k}\big(C_{\hat\alpha i}\mu_i\big) \, \delta^{(4k)}\big(C_{\hat\alpha i}\eta_i\big)} {M_1 M_2 \cdots M_n}. \end{equation} It is here understood that the $C$'s are functions of the $B$'s as given by \reef{DfromC}. Note that the Schouten identity guarantees the following two important properties: \begin{itemize} \item The $C$'s are invariant under the translations \reef{translation}. \item The expression \reef{DfromC} implies that $C_{\hat\alpha i}\lambda_i = 0$. \end{itemize} We would now like to do two things: fix the translational redundancy and rewrite the integral in terms of $C$'s instead of $B$'s. Because of the translational invariance, the $B$'s are not independent variables: for example we can use translations to set $2k$ of them to zero (see below). So after fixing translational invariance, we will have $kn - 2k = k(n-2)$ variables to integrate over. \compactsubsection{Step 1: Fixing translation invariance.} Let us use the translation invariance $T_k$ to fix the first two columns in $B_{\hat\alpha i}$ to be zero, i.e.~for all $\hat{\alpha}=1,\ldots,k$ we set $B_{\hat\alpha 1}=B_{\hat\alpha 2}=0$. This gives \begin{equation} \frac {d^{k\times n} B_{\hat\alpha i}} {T_k} = \<12\>^{k} \,d^{k\times (n-2)} B_{\hat\alpha i}\,, \label{translationfix} \end{equation} where the included prefactor preserves the scaling properties of the measure. (This can be derived more carefully as a Jacobian of the gauge fixing.) \compactsubsection{Step 2: Changing variables from $B$ to $C$} We know how $C_{\hat{\alpha}i}$ is related to $B_{\hat{\alpha}i}$ from equation \eqref{DfromC}. We can use that relation to solve for $k(n-2)$ of the components of $C$ in terms of the $k(n-2)$ unfixed components of $B$. Given our choice to set $B_{\hat{\alpha}1}=B_{\hat{\alpha}2}=0$, we have the following system of $k(n-2)$ equations: \begin{align} C_{\hat{\alpha}{i}} = \left\{ \begin{array}{cll} \displaystyle \frac{\two{i,i+1}B_{\hat\alpha, i-1} + \two{i+1,i-1}B_{\hat\alpha i} + \two{i-1,i}B_{\hat\alpha, i+1}} {\two{i-1,i}\two{i,i+1}} & \text{for}& 3<i<n \\[5mm] \displaystyle \frac{\two{42}B_{\hat\alpha 3} + \two{23}B_{\hat\alpha 4}} {\two{23}\two{34}} & \text{for}& i=3 \\[5mm] \displaystyle \frac{\two{n1}B_{\hat\alpha, n-1} + \two{1,n-1}B_{\hat\alpha n} } {\two{n-1,n}\two{n1}} & \text{for} & i=n \end{array}\right.\,. \label{CfromBgauge12} \end{align} We can write this as $C_{\^\alpha\^j} = B_{\^\alpha\^i}Q_{\^i\^j}$, with $\hat{i},\hat{j} = 3,\ldots, n$, for a square symmetric matrix $Q_{\^i\^j}$ with nonzero entries only on and adjacent to the main diagonal. In Appendix \ref{app:detQ}, we show that \begin{align} |\det Q\,| = \frac{\two{12}^2}{\two{12}\two{23}\cdots \two{n1}}\,.\ \label{detQresult} \end{align} Therefore the measure transforms as \begin{align} d^{k(n-2)} B_{\^\alpha \^i} ~=~ \frac{d^{k(n-2)}C_{\^\alpha\^i}}{|\det Q\,|^k} ~=~ \bigg( \frac{\two{12}\two{23}\cdots \two{n1}} {\two{12}^2} \bigg)^k d^{k(n-2)}C_{\^\alpha\^i}\,. \label{measureBC} \end{align} We take the absolute value of the determinant since the overall sign is irrelevant. Once again, the little group scaling of $dC$ is compensated by the Jacobian factor so that the overall scaling of the right-hand side matches that of $dB$ on the left. \compactsubsection{Step 3: Restoring the full set of $C$ variables} With the help of \reef{translationfix} and \reef{measureBC}, the integral \reef{momgrass2} now takes the form \begin{equation} \label{momgrass3} {\mathcal{L}}_{n;k}\big(\mathcal{Z}\big) = \frac{\delta^{4}\big(\lambda_i \tilde\lambda_i\big) \,\delta^{(8)}\big(\lambda_i \tilde{\eta}_i\big) } {\two{12}\two{23} \cdots \two{1n}} \frac{1}{\<12\>^k} \int \frac{d^{k\times (n-2)} C_{\hat\alpha i}} {GL(k)} \, \frac{\delta^{2k}\big(C_{\hat\alpha i}\mu_i\big) \, \delta^{(4k)}\big(C_{\hat\alpha i}\eta_i\big)} {M_1 M_2 \cdots M_n} \bigg|_{C_{\hat\alpha 1,2} = C_{\hat\alpha 1,2}^{(0)}} \,. \end{equation} We recognize the first factor as the MHV superamplitude $\mathcal{A}_n^\text{MHV}$ from \reef{AMHV}. The restriction of $C_{\hat\alpha 1}$ and $C_{\hat\alpha 2}$ in \reef{momgrass3} follows from the relation \reef{DfromC} between the $B$ and $C$; it was used above to solve for $k(n-2)$ components of $C$ in terms of the $B$'s, but the remaining $2k$ components of $C$ are then fixed as \begin{equation} C_{\^\alpha 1}= \frac{B_{\^\alpha n}}{\two{n1}} = \sum_{j=3}^n \frac{\two{2j}}{\two{12}} C_{\^\alpha j} =: C_{\^\alpha 1}^{(0)}\,, ~~~~~~ C_{\^\alpha 2}= \frac{B_{\^\alpha 3}}{\two{23}} = -\sum_{j=3}^n \frac{\two{1j}}{\two{12}} C_{\^\alpha j} =: C_{\^\alpha 2}^{(0)}\,. \label{C1C2} \end{equation} To verify the second equality in each relation, use \reef{CfromBgauge12} and rejoice in the beauty of the sum telescoping under the Schouten identity. Thus, when evaluating the integral \reef{momgrass3}, $C_{\hat\alpha 1}$ and $C_{\hat\alpha 2}$ are functions of the other $k(n-2)$ $C$-components. This is a restriction of the region of integration that we can also impose via $2k$ delta functions $\delta \big( C_{\^\alpha i} - C_{\^\alpha i}^{(0)} \big)$ for $i=1,2$. Moreover, it follows from the explicit solution \reef{C1C2} that the constraints are equivalent to $C_{\^\alpha i}\lambda_i = 0$. Thus we can rewrite the delta function restriction as \begin{equation} \delta \big( C_{\^\alpha 1} - C_{\^\alpha 1}^{(0)} \big)\, \delta \big( C_{\^\alpha 2} - C_{\^\alpha 2}^{(0)} \big) = \<12\> \,\delta^2 \big( C_{\^\alpha i}\lambda_i\big)\,. \end{equation} for each $\hat{\alpha}=1,2,\ldots,k$. We then have \begin{equation} \label{momgrass4} {\mathcal{L}}_{n;k}\big(\mathcal{Z}\big) = \mathcal{A}_n^\text{MHV} \int \frac{d^{k\times n} C_{\hat\alpha i}} {GL(k)} \, \frac{\delta^{2k} \big( C_{\^\alpha i}\lambda_i\big)\, \delta^{2k}\big(C_{\hat\alpha i}\mu_i\big) \, \delta^{(4k)}\big(C_{\hat\alpha i}\eta_i\big)} {M_1 M_2 \cdots M_n} \,. \end{equation} Although we chose to fix the first two columns of $B$ to be zero, the answer is independent of that choice; the factors of $\two{12}$ cancel out. We can now write the result directly in terms of the momentum supertwistors $\mathcal{Z}_i=( Z_i | \eta_i )= ( \lambda_i, \mu_i | \eta_i )$ as \begin{equation} \label{momgrass5} {\mathcal{L}}_{n;k}\big(\mathcal{Z}\big) = \mathcal{A}_n^\text{MHV} \int \frac{d^{k\times n} C_{\hat\alpha i}} {GL(k)} \, \frac{\delta^{4k} \big( C_{\^\alpha i} Z_i\big)\, \delta^{(4k)}\big(C_{\hat\alpha i}\eta_i\big)} {M_1 M_2 \cdots M_n} = \mathcal{A}_n^\text{MHV} \int \frac{d^{k\times n} C_{\hat\alpha i}} {GL(k)} \, \frac{\delta^{4k|4k} \big( C_{\^\alpha i} \mathcal{Z}_i\big)} {M_1 M_2 \cdots M_n} \,. \end{equation} This completes our derivation of the $\mathcal{N}=4$ SYM Grassmannian integral in momentum twistor space from that in momentum space. A very similar procedure leads to an analogous result in 3d ABJM theory as we explain below in Section \ref{sec:3DGrass}. In the intervening section, we demonstrate an explicit evaluation of the ${\cal N}=4$ SYM momentum twistor integral \eqref{momgrass5} in the NMHV sector. \section{NMHV Integrals and Residues} \label{s:res} In this section we evaluate the NMHV Grassmannian integral in momentum twistor space and discuss some properties of the residues and their relations to on-shell diagrams. While part of this is review, new material includes recasting the residue theorems in terms of the homology and a precise description of how the residue relations and pole structures relate to the boundary operation and the boundaries of cells in the Grassmannian. \subsection{Evaluation of the NMHV residues} \label{s:NMHVeval} We focus on the momentum twistor Grassmannian, so for NMHV we have $k=1$ and \reef{GrassZ} is a contour integral in the Grassmannian $\text{G}(1,n)$. The elements $C\in \text{G}(1,n)$ are $1\times n$ matrices modulo a $GL(1)$ scaling, \begin{align} C = \big[\begin{array}{cccc} c_1 & c_2 & \ldots & c_n\end{array}\big], \label{topC} \end{align} with complex numbers $c_i$. The Grassmannian integral is $\mathcal{L}_{n;1}(\mathcal{Z}) = \mathcal{A}_n^\text{MHV} \mathcal{I}_{n;1}(\mathcal{Z})$ with \begin{align} \mathcal{I}^{(\Gamma)}_{n;1}(\mathcal{Z}) := \oint_\Gamma \frac{d^{1\times n}C}{GL(1)\, c_1 c_2 \cdots c_n} \,\delta^4\big(c_i Z_i\big)\, \delta^{(4)}\big(c_i {\eta}_i\big)\,. \label{NMHVlink} \end{align} The oriented volume form on $\mathbb{C}^n$ is $d^nC = \bigwedge\limits_{i=1}^n\,dc^i$, and the contour $\Gamma$ will be specified below. The bosonic delta function $\delta^4(c_i Z_i)$ fixes four $c_i$'s, and the $GL(1)$ redundancy fixes another. This leaves an integral with $n-5$ variables. Now suppose the contour $\Gamma$ encircles a pole where exactly $n-5$ of the $c_i$'s vanish. Such a contour can be characterized by specifying which five $c_i$'s are non-vanishing at the pole. Let us denote these five non-vanishing $c_i$'s by $c_a$, $c_b$, $c_c$, $c_d$, and $c_e$, and the corresponding contour $\gamma_{abcde}$. We now evaluate $\mathcal{I}_{n;1}^{\gamma_{abcde}}(\mathcal{Z})$ ``by inspection". Appendix \ref{app:residues} gives a more careful evaluation that also computes the sign of the residue correctly. It follows from \reef{NMHVlink} that the residue where all $c_i$ vanish for $i\ne a,b,c,d,e$ is, up to a sign, simply \begin{equation} \mathcal{I}_{n;1}^{\gamma_{abcde}}(\mathcal{Z}) = \frac{\delta^4\big(c_a Z_a + c_b Z_b + c_c Z_c+ c_d Z_d + c_e Z_e\big)\, \delta^{(4)}\big(c_a \eta_a + c_b \eta_b + c_c \eta_c+ c_d \eta_d + c_e \eta_e \big)}{c_a c_b c_c c_d c_e}\,. \label{Ires} \end{equation} Now, the constraint enforced by the bosonic delta-function is trivially solved by \begin{equation} c_a = \<bcde\>\,,~~~~ c_b = \<cdea\>\,,~~~~ c_c = \<deab\>\,,~~~~ c_d = \<eabc\>\,,~~~~ c_e = \<abcd\> \,, \end{equation} using the 5-term Schouten identity (or Cramer's rule) that states that five 4-component vectors are necessarily linearly dependent: \begin{align} \four{ijkl}Z_m +\four{jklm}Z_i + \four{klmi}Z_j + \four{lmij}Z_k + \four{mijk}Z_l=0 \,. \label{5Schouten} \end{align} The 4-brackets are the fully antisymmetric $SU(2,2)$-invariants \begin{equation} \label{def4bracket} \<ijkl\> := -\epsilon_{\mathsf{ABCD}} Z_i^\mathsf{A}Z_j^\mathsf{B}Z_k^\mathsf{C}Z_l^\mathsf{D} = \det\big( Z_i Z_j Z_k Z_l\big) \,. \end{equation} We conclude that \begin{equation} \mathcal{I}_{n;1}^{\gamma_{abcde}} = \frac{\delta^{(4)}\Big(\four{bcde}{\eta}_a + \four{cdea}{\eta}_b+\four{deab}{\eta}_c+\four{eabc}{\eta}_d + \four{abcd}{\eta}_e\Big)}{\four{bcde} \four{cdea} \four{deab} \four{eabc} \four{abcd} } ~=: \five{abcde}\,, \label{resresult} \end{equation} The expression \reef{resresult} is manifestly antisymmetric in the five labels $a,b,c,d,e$. This follows from the standard evaluation of higher-dimensional contour integrals, as we review in Appendix \ref{app:residues}. In addition, the general results in the appendix tell us that the residue is also fully antisymmetric in the labels $i\ne a,b,c,d,e$. We can incorporate that by labeling the residue \eqref{Ires}, including the appropriate signs from the appendix, by the $n-5$ values $i_1,i_2,\ldots,i_{n-5} \ne a,b,c,d,e$ as $\{i_1,i_2,\ldots,i_{n-5}\}$, which is antisymmetric in its indices. Then the final answer, which includes all of the signs from Appendix \ref{app:residues}, is \begin{align} \mathcal{I}_{n;1}^{\gamma_{abcde}} = \frac{1}{(n-5)!}\, \varepsilon^{ a\,b\,c\,d\,e\,i_1\, i_2\,\ldots\, i_{n-5}}\ \{i_1,i_2,\ldots,i_{n-5}\} = \five{abcde} \label{5res} \,. \end{align} This completes the calculation of the residues of the Grassmannian integral $\mathcal{L}_{n;1}$ in momentum twistor space. The result, \begin{equation} \mathcal{L}_{n;1}^{\gamma_{abcde}} = \mathcal{A}_n^\text{MHV} \,\five{abcde} \,, \end{equation} shows that the individual residues produced by $\mathcal{L}_{n;1}$ are the 5-brackets $\five{abcde}$. These are the known building blocks of NMHV amplitudes, both at tree and loop-level. \subsection{NMHV residue theorems} \label{s:resrels} Since the residues $\five{abcde}$ of the NMHV Grassmannian integral \reef{NMHVint} are characterized by five labels, $a,b,c,d,e \in \{1,2,3,\ldots,n\}$, as in \reef{5res}, it follows that there are a total of ${n \choose 5}$ NMHV residues. These, however, are not independent. While it is difficult to derive the residue relations --- or even verify them --- by direct computations, the constraints among them follow quite straightforwardly from the Grassmannian residue theorems, as first noted in \cite{ArkaniHamed:2009dn}. In this section, we count the number of independent NMHV residues $\five{abcde}$ and examine the linear relations among them. Since the only input is residue theorems, these relationships are also true off the support of the external momentum and supermomentum delta functions in the overall MHV factor in the momentum twistor Grassmannian integral. Let us begin by taking an abstract view of the integral \reef{NMHVint}. We are integrating over $C = \big[ c_1 \ldots c_n \big]$ modulo a $GL(1)$ that identifies $C \sim s C$ for any $s \in \mathbb{C}-\{0\}$. Thus $C$ can be viewed as homogeneous coordinates of $\mathbb{CP}^{n-1}$. We are interested in the residues associated with simultaneously vanishing `minors' $c_i$. Each condition $c_i=0$ defines a hyperplane in $\mathbb{CP}^{n-1}$. In other words, we are interested in the $n$ hyperplanes $h_i := \{ C \in \mathbb{CP}^{n-1} | c_i = 0 \}$. This is called a {\em hyperplane arrangement} in $\mathbb{CP}^{n-1}$. Specifically, the residue $\five{abcde}$ corresponds to picking up the residue from the $(n-5)$-dimensional toroidal contour $(S^1)^{n-5}$ surrounding the intersection of the $n-5$ hyperplanes $h_i$ with $i \ne a,b,c,d,e$. The bosonic delta functions in the momentum twistor integral \reef{NMHVint} impose four conditions among the $n$ components of $C$. These homogeneous linear relations respect the $GL(1)$ scaling, so they reduce the space of interest from $\mathbb{CP}^{n-1}$ to $\mathbb{CP}^{n-5}$. Consequently, we are interested in the arrangement of $n$ hyperplanes in $\mathbb{CP}^{n-5}$. Now, suppose we focus on the complement of $h_n$, i.e. $c_n\neq0$. We fix $c_n$ to be some non-vanishing value to eliminate the projective freedom, so $\mathbb{CP}^{n-5}\rightarrow \mathbb{C}^{n-5}$. The problem then reduces to the study of $n-1$ hyperplanes $\{h_i\}_{i=1,\ldots,n-1}$ in $\mathbb{C}^{n-5}$. This step is equivalent to fixing the $GL(1)$ redundancy in the Grassmannian integral. The $(n-5)$-dimensional contours of \eqref{NMHVint} must therefore live in the hyperplane arrangement complement \begin{equation} X= \mathbb{C}^{n-5} - \bigcup_{i<n} h_i\,. \end{equation} A residue does not change under continuous deformation of the contour, so the result only depends on the homology class of the contour. Thus, the key observation is that the number of possible independent residues is the dimension of the homology class $H_{n-5}(X,\mathbb{C})$. The geometry of hyperplane arrangements is well-studied in the mathematics literature and the results include the following theorem \cite{OrlikTerao}: \begin{theorem} Let $X = \mathbb{C}^N-\bigcup\limits_{i=1}^r h_i$ be a hyperplane arrangement complement. Then \begin{enumerate} \item The cohomology $H^*(X,\mathbb{C})$ is generated by the forms $\frac{d\alpha_i}{\alpha_i}$. \item Suppose $\{h_i\}$ is generic.\footnote{When the ambient space is $\mathbb{C}^N$, we say that the hyperplane arrangement is {\it generic} if $h_{i_1} \cap h_{i_2} \cap \cdots \cap h_{i_r}$ has dimension $N-r$, for $r \leq n$. In other words, a hyperplane arrangement is generic if all intersections of hyperplanes have the expected dimension.} Then $H^k(X,\mathbb{C})$ is zero for $k \geq N$, and for $k \in \{0,1,\ldots,N\}$, a basis for $H^k(X,\mathbb{C})$ is given by the forms $$ \frac{d\alpha_{i_1}}{\alpha_{i_1}} \wedge \cdots \wedge \frac{d\alpha_{i_k}}{\alpha_{i_k}} $$ ranging over subsets $\{i_1,i_2,\ldots,i_k\} \subset \{1,2,\ldots,r\}$. In particular, \begin{equation} \dim H^k(X,\mathbb{C}) = {{r}\choose{k}}\,. \end{equation} \end{enumerate} \end{theorem} The algebra $H^*(X,\mathbb{C})$ (generic arrangement or otherwise) can be described in a combinatorial fashion and is called the Orlik-Solomon algebra. For our purpose, the ambient space has dimension $N=n-5$, there are $r=n-1$ hyperplanes, and we are interested in the dimension of the homology $H_{n-5}(X,\mathbb{C})$. It is of course the same dimension as the corresponding cohomology $H^{n-5}(X,\mathbb{C})$. Hence, by the above theorem, the number of independent residues is \begin{align} R=\dim H_{n-5}(X) = {{n-1}\choose{n-5}} = {{n-1}\choose{4}} \,. \label{indepRes} \end{align} The residue relations have a simple geometrical interpretation (see also \cite{ArkaniHamed:2009dn}). Let $\{i_1,i_2,\ldots,i_{n-5}\}$ denote the residue corresponding to the intersection of $n-5$ hyperplanes ${h_{i_1} \cap h_{i_2} \cap \cdots \cap h_{i_{n-5}}}$. As explained in the Section \ref{s:NMHVeval}, the residue is fully antisymmetric in its labels. For example, when $n=6$, the ``hyperplanes'' are just individual points $\{i\}$ in $\mathbb{CP}^1$; there are six such points. Since $\mathbb{CP}^1$ is isomorphic to a two-sphere $S^2$, any contour surrounding all six can be contracted to a point. Hence the residue theorem states that the sum of the six residues is zero. Thus there is one relation among six residues, leaving five independent. This clearly agrees with the counting \reef{indepRes} for $n=6$. Let us now use this to understand the relations under which only $n-1 \choose 4$ of the $n \choose 5$ residues are independent. Consider a choice of $n-6$ hyperplanes, $h_{i_k}$ with $k=1,2,\ldots,n-6$, in $\mathbb{CP}^{n-5}$. Imagine that we take the $S^1$ contours surrounding each of these $h_i$ very small so that we effectively look at the subspace $\mathbb{CP}^1 = S^2$ of the intersection of those $n-6$ hyperplanes. This subspace is (generically) intersected by the other hyperplanes $h_j$ at $6$ distinct points. Just as for the $n=6$ case, a contour in $\mathbb{CP}^1$ that surrounds these six points can be contracted a point, and the sum of the six residues must vanish: the resulting residue theorem is \begin{align} \sum_{j=1}^n \{i_1,i_2,\ldots,i_{n-6},j\} = 0 \,. \label{resCon} \end{align} This holds for any choice of $n-6$ labels $i_1,i_2,\ldots,i_{n-6}$; hence we get a web of linear relations among the $n \choose 5$ residues. The statement \reef{indepRes} is that under these relations, only $n-1 \choose 4$ residues are independent. The counting of independent residues can be verified directly from the relations \reef{resCon}. While there may appear to be $n \choose {n-6}$ constraints in \reef{resCon}, some of them are redundant. Without loss of generality, consider only those for which all $i_k\neq n$, $k=1,2,\ldots,n-6$. There are ${n-1} \choose {n-6}$ distinct constraints of that sort. In each such sum, the index $n$ will appear exactly once, namely when $j=n$, so we can solve for each residue that includes $n$ in terms of residues which do not: \begin{align} \{i_1,i_2,\ldots, i_{n-6},n\} = - \sum\limits_{j=1}^{n-1}\{i_1,i_2,\ldots,i_{n-6},j\} \label{nthRes} \end{align} where $i_1,i_2,\ldots, i_{n-6} \ne n$. This determines all of the residues labeled by $n$ in terms of all of the others. Furthermore, since the first $n-6$ indices form a unique set, all ${n-1}\choose{n-6}$ equations in \eqref{nthRes} are independent. The remaining equations in \eqref{resCon} have $i_k =n$ for some $k$, but they do not provide any further constraints. To see this, use the antisymmetry and \eqref{nthRes} to eliminate the index $n$: \begin{align} \sum\limits_{j=1}^{n}\{i_1,i_2,\ldots,i_{n-7},n,j\}&=-\sum\limits_{j=1}^{n}\sum\limits_{m=1}^{n-1}\{i_1,i_2,\ldots,i_{n-7},m,j\} \nonumber\\ &=-\sum\limits_{m=1}^{n-1} \Big(\sum\limits_{j=1}^{n}\{i_1,i_2,\ldots,i_{n-7},m,j\}\Big)=0\,. \end{align} We conclude that all of the constraints with a fixed index $n$ are redundant with the ones in \reef{nthRes}. Hence, the number of independent constraints are ${{n-1}\choose{n-6}}$ and therefore the number of independent residues is \begin{align} R = {{n}\choose{n-5}} - {{n-1}\choose{n-6}} = {{n-1}\choose{n-5}} = {{n-1}\choose{4}}\,, \end{align} in agreement with the dimension of the homology \reef{indepRes}. \subsection{Applications} \label{s:resapp} \subsubsection{Residue theorems as boundary operations} Let us now consider some applications of the NMHV residue theorems. The case of $n=6$ is very well-known. There is just one constraint from \reef{resCon}, \begin{align} \{1\}+\{2\}+\{3\}+\{4\}+\{5\}+\{6\}=0\,, \end{align} or via \reef{5res} in terms of the 5-brackets it is the six-term identity \begin{equation} \five{23456} - \five{13456} + \five{12456} - \five{12356} + \five{12346} - \five{12345} = 0\,. \end{equation} The LHS of this identity can be succinctly abbreviated as defining the {\bf \em boundary operation} $\partial \big[123456\big]$; the relation to boundaries is explain in Section \ref{s:bdr}. More generally, we can write the boundary operation as \begin{equation} \partial \big[abcdef\big] = 0 ~~~~\longleftrightarrow~~~~ \sum_{a',b',c',d',e',f'=1}^n \epsilon^{i_1 i_2 \ldots i_{n-6} a'b'c'd'e'f'} [a'b'c'd'e'] = 0\,, \label{bdrOp} \end{equation} where $\{ i_1, \ldots ,i_{n-6} \}$ are the complement of $\{a,b,c,d,e,f\}$ in the set $\{1,2,\ldots,n\}$. The relation \reef{5res} between the 5-brackets and the residues now makes it clear that the boundary conditions are equivalent to the residue theorems \reef{resCon}: \begin{equation} \sum_{j=1}^n \{i_1,i_2,\ldots,i_{n-6},j\} = 0 ~~~~ \raisebox{3mm}{$\underleftrightarrow{~~ \scriptstyle \{i_1,i_2,\ldots,i_{n-6}\}\; = \; \overline{\{a,b,c,d,e,f\}}~~}$} ~~~~ \partial\big[abcdef\big] = 0\,. \end{equation} \subsubsection{Identities among $R$-invariants} Prior to the introduction of momentum twistors, the momentum space versions of the 5-brackets were denoted as {\bf \em $R$-invariants} \cite{Drummond:2008vq}: \begin{equation} R_{ijk} := \five{i,j-1,j,k-1,k}\,. \end{equation} It was observed that the $R$-invariants obey the two identities \begin{equation} R_{i,i+2,j} = R_{i+2,j,i+1}\, ~~~~\text{and}~~~~ \sum_{s=3}^{k-2} \sum_{t=s+2}^k R_{1st} = \sum_{s=2}^{k-3} \sum_{t=s+2}^{k-1} R_{kst} \,. \label{Ridentities} \end{equation} for any $k=1,2, \dots, n$. These identities have been used in various applications, such as proving dual conformal invariance of the 1-loop ratio function in ${\cal N}=4$ SYM \cite{Drummond:2008vq,Brandhuber:2009xz,Elvang:2009ya}. Let us now review how these arise as a consequence of the symmetries and residue theorems of the 5-brackets. The first identity in \reef{Ridentities} follows straightforwardly from the antisymmetry of the 5-bracket \cite{Mason:2009qx}: \begin{equation} R_{i,i+2,j} = \five{i,i+1,i+2,j-1,j} =\five{i+2,j-1,j,i,i+1} = R_{i+2,j,i+1} \,. \end{equation} For the second identity in \reef{Ridentities}, note that $R_{1st} = \five{1,s-1,s,t-1,t}$ vanishes for $s=2$, so on the LHS of \reef{Ridentities} the sum can trivially be extended to include $s=2$. Then using the six-term identity resulting from $\partial \five{1,s-1,s,t-1,t,k} =0$, the 5-bracket $\five{1,s-1,s,t-1,t}$ can be eliminated in favor of the five other 5-brackets appearing in the identity. This includes $\five{s-1,s,t-1,t,k} = \five{k,s-1,s,t-1,t}$, which vanishes trivially for $t=k$ and for $s=k-2$. Hence this part of the sum gives the desired RHS of \reef{Ridentities}. We are left to show that the sum of the remaining four terms vanishes; they are \begin{equation} \sum_{s=2}^{k-2} \sum_{t=s+2}^k \Big( \five{1,s-1,s,t-1,k} - \five{1,s-1,s,t,k} +\five{1,s-1,t-1,t,k}-\five{1,s,t-1,t,k} \Big)\,. \label{left4terms} \end{equation} The sum of the first two terms telescopes to $\sum_{s=3}^{k-2} \five{1,s-1,s,s+1,k}$ while the sum of the last two terms collapses to $-\sum_{s=2}^{k-3} \five{1,s,s+1,s+2,k}$. These two sums are identical and thus the sum \reef{left4terms} vanishes. This completes the derivation of the identities \reef{Ridentities}. \subsubsection{Locality and the NMHV tree superamplitude} \label{s:tree} For $k=n$, the second identity in \reef{Ridentities} can be written \begin{equation} \sum_{i<j} \five{1,i-1,i,j-1,j} = \sum_{i<j} \five{n,i-1,i,j-1,j}\,. \end{equation} Note that in this representation, the first label on the LHS plays no special role and can be replaced with any momentum twistor $Z_*$. Hence the sum of $\five{*,i-1,i,j-1,j}$ over all $i<j$ is independent of $Z_*$. Let us now study the {\bf \em pole structure} of the 5-brackets. A given 5-bracket has five poles, namely where each of the five 4-brackets in the denominator vanish. Consider two 5-brackets that differ by just one momentum twistor, e.g.~$\five{abcdx}$ and $\five{abcdy}$. They share one common pole, namely $\four{abcd}$. The singularity occurs on the subspace where the four momentum twistors $Z_{a,b,c,d}$ become linearly dependent. Since $Z_y \in \mathbb{CP}^3$ can be expressed as a linear combination of any four other (linearly independent) momentum twistors, we can write $Z_y = w_x Z_x + w_a Z_a + w_b Z_b + w_c Z_c$. Using this, it is straightforward to show that the residue at the pole $\four{abcd}=0$ is the same for $\five{abcdx}$ and $\five{abcdy}$. In other words, the residue of the pole $\four{abcd}$ vanishes in the combination $\five{abcdx}-\five{abcdy}$. It is natural to associate a boundary operation with the residues of the poles of the 5-brackets, written as \begin{equation} \label{bdr5bracket} \partial \five{abcde} := \five{bcde} - \five{acde}+\five{abde}-\five{abce}+\five{abcd} \,. \end{equation} The signs keep track of the relative signs of the residues. It now follows that the cancellation of $\five{abcd}$ in $\partial \big( \five{abcdx}-\five{abcdy} \big)$ is equivalent to the statement that the residue of the pole at $\four{abcd}=0$ vanishes in the difference of the two five-brackets. Physical poles in color-ordered tree-level scattering amplitudes are exactly those associated with vanishing Mandelstam invariants $(p_i + p_{i+1} + \ldots)^2$ involving a sum of a subset of adjacent momenta. These are precisely associated with poles in the 5-brackets of the form $\four{i-1,i,j-1,j}$ because of the identity \cite{Hodges:2009hk,Mason:2009qx} \begin{equation} \big(p_i + p_{i+1} + \ldots p_{j-1}\big)^2 = \frac{\four{i-1,i,j-1,j}}{\<i-1,i\>\<j-1,j\>}\,. \end{equation} Poles \emph{not} of the form $\four{i-1,i,j-1,j}$ are spurious: they cannot appear in the tree-amplitude. A straightforward algebraic exercise shows that \begin{equation} \partial \sum_{i<j} \five{*,i-1,i,j-j,j} = \sum_{i<j} \five{i-1,i,j-j,j} \,. \end{equation} This means that all spurious poles in the LHS sum telescope to zero, leaving just the manifestly local poles. For $* = q = 1,2,\dots,n$, this sum --- times the MHV superamplitude --- is exactly the expression one finds \cite{Drummond:2008cr} as the solution to the BCFW recursion relation based on a $[q,q+1\>$-supershift of the {\bf \em tree NMHV superamplitude} \begin{equation} \mathcal{A}_n^\text{NMHV} = \mathcal{A}_n^\text{MHV} \sum_{i<j} \five{*,i-1,i,j-j,j} \,, \label{NMHVtree} \end{equation} for $Z_* = Z_q$.\footnote{When $Z_*$ is not selected to be one of the $n$ momentum twistors $Z_i$ of the external data, the expression \reef{NMHVtree} is a CSW-like representation of the NMHV superamplitude.} From the point of view of the Grassmannian, we see that \reef{NMHVtree} results from a certain choice of contour. Thus, there are choices of contours for the Grassmannian integral such that the result is exactly the NMHV tree amplitude; such a contour what we called the `tree contour'. Note that the insistence of locality, in the sense of having only physical poles, allowed us to identify the tree contours. It may seem puzzling that only a small subset of the residues produced by the Grassmannian integral appear in the BCFW-form \reef{NMHVtree} of the NMHV tree superamplitude: residues of the form $\five{q,i-1,i,j-1,j}$ are used, while residues such as $\five{1,2,4,6,8}$ or $\five{1,3,5,7,9}$ do not seem to play a role, other than through the residue theorems. It would be peculiar if the other residues of the same Grassmann degree were not relevant for NMHV amplitudes; it turns out that they are. It has been conjectured \cite{ArkaniHamed:2009dn} that --- in addition to the tree superamplitudes --- the Grassmannian integral also produces all the Leading Singularities of all amplitudes in planar ${\cal N}=4$ SYM at any loop order. The 1-loop NMHV ratio function \cite{Drummond:2008vq,Brandhuber:2009xz,Elvang:2009ya} can be written in terms of the exactly the same types of residues $\five{q,i-1,i,j-1,j}$ as at tree-level, so one has to go to 2-loop order to encounter `non-tree' residues in the Leading Singularities \cite{ArkaniHamed:2009dn}. Also, it has been demonstrated that no new Leading Singularities appear beyond 3-loop order in the NMHV sector \cite{Bullimore:2009cb,Broedel:2010rr}, so the first three loop-orders of the NMHV amplitudes are expected to utilize the full set of residues produced by the Grassmannian integral $\mathcal{L}_{n;1}$. \subsection{Cells, permutations, and on-shell diagrams} \label{s:onshelldiag} So far we have described the evaluation of NMHV amplitudes in the language of contour integrals and residue theorems, but we find that it is also instructive to take a more abstract view and consider how it fits into the context of on-shell diagrams, permutations, and cells of the Grassmannian. Since the calculations in Section \ref{s:res} were performed in the momentum twistor formulation, we will discuss that case first, and subsequently develop the corresponding story in the momentum space formulation. Before delving into the details, it will be helpful to quickly review some terminology from \cite{ArkaniHamed:2012nw}. Subspaces of the Grassmannian G$(k,n)$ can be classified into {\bf \em cells} by specifying the ranks of cyclically consecutive columns $C_i$, that is ${\rm rank}\big({\rm span}(C_i,C_{i+1},\ldots,C_j)\big)$ for all cyclic intervals $[i,j]$. The {\em dimension} of a cell is the number of parameters it takes to specify a matrix representative modulo the $GL(k)$ redundancy. Cells are uniquely labeled by decorated permutations, which are ``permutations" of the set $\{1,2,\ldots,n\}$ in which $k$ of the elements are shifted beyond $n$.\footnote{The decorated permutations will be familiar to practitioners of the juggling arts, where they are also referred to as ``juggling patterns." The decoration encodes that balls can only be thrown forward.} Throughout the remainder of this text, we will use `permutation' and `decorated permutation' interchangeably, but we will always mean the latter. Each permutation labels a cell by encoding the linear dependencies of the columns in a representative matrix of the cell. Treating the matrix columns $c_i$ as $k$-vectors, a given permutation \begin{align} \sigma = \{\s(1),\s(2),\ldots\} = \{a,b,\ldots\} \end{align} encodes that $c_a$ is the first column with $a>1~($mod $n)$ such that $c_1$ is in the span of $\{c_2,c_3,\ldots,c_a\}$. Similarly, $c_2$ is spanned by $\{c_3,\ldots,c_b\}$, and so on. Entries for which $\sigma(i)=i$ imply that the $i^\text{th}$ column is identically zero. As an example, consider the 5-dimensional cell in G$(2,6)$ with representative matrix \begin{align} \left( \begin{array}{cccccc} 1 & 0 & c_{11} & c_{12} & 0 & c_{14} \\ 0 & 1 & c_{21} & c_{22} & 0 & 0 \\ \end{array} \right)\,. \end{align} One can easily verify that this cell is labeled by the permutation $\sigma=\{3,4,6,8,5,7\}$. The cell with maximal dimension $k(n-k)$ in G$(k,n)$ is known as the {\bf \em top cell}, and it is the unique cell in which at a generic point none of the $k\times k$ minors vanish in a representative matrix. Since none of the consecutive minors vanish, each column must be spanned by the next $k$ columns, and therefore the top cell is labeled by a permutation of the form \begin{align} \sigma_\text{top} = \{1+k,2+k,\ldots,n+k\}\,. \end{align} \subsubsection{Momentum twistor space} The $n$-particle NMHV integral in momentum twistor space \eqref{NMHVlink} is an integral over the $(n-1)$-dimensional top cell of G$(1,n)$, which has a representative $1\times n$ matrix $C$, as in \eqref{topC}. Since the minors of $C$ are determinants of $1\times1$ matrices (i.e.~numbers), the top cell is represented by matrices with all non-zero entries. The external data enters through the delta functions $\delta^{4|4}(C\cdot\mathcal{Z})$. The four independent bosonic delta functions fix all degrees of freedom for any 4-dimensional cell of G$(1,n)$. In order to reach a 4d cell from the top cell, one must set $n-5$ of the minors to zero. This is done in practice by choosing an $(n-5)$-dimensional contour $\gamma_{abcde}$ that encircles a point where only five of the coordinates, i.e. minors, are non-vanishing. For a given choice of $a,b,c,d,e$, the result of evaluating the contour integral is an integral (that will be fully localized by the delta functions) over a unique 4d cell labeled by the decorated permutation \begin{align} & {\color{blue}\,a ~~~~~~~~~~~~~~b ~~~~~~~ c ~~~~~~~ d ~~~~~~~~ e} \nonumber\\[-2mm] \sigma_{abcde} = \{1,2,\ldots,a-1,\, & b,a+1,\ldots,c,\ldots,d,\ldots,e,\ldots,a+n,\ldots,n\}\,, \end{align} where all entries are self-identified except those at positions $a,b,c,d,e$ (marked above). The bosonic delta functions fix the remaining degrees of freedom and leave the residue $\{i_1,i_2,\ldots,i_{n-5}\}$, which is related to the five-bracket $\five{abcde}$ via equation \eqref{5res}. In the momentum twistor Grassmannian integral, all 4d cells meet the support of the delta-function at NMHV level thanks to the 5-term Schouten identity \reef{5Schouten}. (This is specific to NMHV level; it is not the case for higher $k$ amplitudes.) \subsubsection{Momentum space} The momentum space version of the Grassmannian integral was given in \reef{GrassM}. For NMHV we have $\tilde{k}=k+2=3$, so the relevant Grassmannian is G$(3,n)$. The top cell of G$(3,n)$ is $(3n-9)$-dimensional and is labeled by the decorated permutation \begin{align} \widetilde{\sigma}_\text{top} = \{4,5,\ldots,n+3\}\,. \end{align} The representative matrices are $3\times n$ and for the top cell all $3\times3$ minors are non-vanishing. The momentum space Grassmannian integral \reef{GrassM} has $2n$ bosonic delta functions, and as we noted in Section \ref{s:grassmMom}, four of them ensure external momentum conservation; similarly, eight of the fermionic delta functions impose supermomentum conservation. The remaining bosonic delta functions fix all degrees of freedom in $(2n-4)$-dimensional cells of G$(3,n)$. Each of these cells is labeled by a unique decorated permutation of the appropriate dimension. Each permutation is also associated with a representative {\bf \em on-shell diagram} (or plabic graph) following the techniques introduced in \cite{ArkaniHamed:2012nw}. For example, the permutation $\widetilde{\sigma}=\{3,5,6,7,9,8,11\}$ is represented by the graph \begin{equation} \raisebox{-1.5cm}{\includegraphics[height=3cm]{red-7pt-35679811-LRpath}}~\,. \label{fig:osd} \end{equation} The permutation is obtained from the graph by following the `left-right paths' from each external leg, turning left at white vertices and right at black vertices; the figure \eqref{fig:osd} shows one such path, yielding $\widetilde{\sigma}(4)=7$. The value of each on-shell diagram is computed by associating each black/white vertex with a 3-point superamplitude, MHV or anti-MHV, respectively. External lines carry the information of the external data while for each internal line an integral must be performed over the corresponding momentum and Grassmann-variables. For details, see \cite{ArkaniHamed:2012nw}. The point that will be important for us in the following is that the vertices enforce special 3-particle kinematics, namely a white vertex with legs has $\lambda_a \propto \lambda_b \propto \lambda_c$ while a black vertex imposes the equivalent condition on the $\tilde{\lambda}$'s. In order to go from the $(3n-9)$-dimensional top cell in G$(3,n)$ to a $(2n-4)$-dimensional cell, we need to eliminate $(3n-9)-(2n-4)=n-5$ degrees of freedom by taking an $(n-5)$-dimensional contour around singularities of the integrand. Of course one could evaluate the integral by first changing variables following the procedure of Section \ref{s:TandMT} and treating the resulting integral in momentum twistor space, but we would like to treat the integral directly in momentum space. Unfortunately, this turns out to be a difficult problem due to the non-linear nature of the $n$ consecutive minors in the denominator. It is not \emph{a priori} sufficient to simply take $n-5$ of the minors to vanish. While this would land in a cell of the correct dimension, there are many $(2n-4)$-dimensional cells which cannot be reached by this technique. This failure is due to the appearance of so-called `composite singularities', which occur when one or more of the minors factorize on the zero-locus of a subset of the coordinates \cite{ArkaniHamed:2009dn}. The dlog forms constructed in \cite{ArkaniHamed:2012nw} resolve many of these difficulties, but there are still some challenges associated with such forms as we mention in the Outlook. \subsubsection{Relating the spaces} In general, there are more $(2n-4)$-dimensional cells in G$(3,n)$ than $4$-dimensional cells in G$(1,n)$, so it may seem puzzling at first that the twistor and momentum twistor integrals of Section \ref{s:TandMT} are supposed to be equivalent. However, at this point we have not yet imposed the bosonic delta functions. NMHV cells are special because all 4d momentum twistor cells meet the support of the delta function. The same is not true of the momentum space cells; all, and only those, cells which meet the delta function support also have momentum twistor duals, but there are some momentum space cells of the correct dimension ($2n-4$) which do not intersect the delta function support and therefore have no corresponding cell in the Grassmannian with momentum twistor formulation. From the associated on-shell diagrams it is easy to see which NMHV momentum space cells will not be supported for generic momenta: any two external legs that are connected by a path containing vertices of only one color are forced by the delta functions at each vertex to have parallel momenta. This condition is not satisfied for generic external data; hence those residues vanish. For example, consider the 10d cell in G$(3,7)$ labeled by the permutation $\widetilde\sigma = \{4,5,6,7,9,10,8\}$, which is shown below \begin{equation} \centering \raisebox{-0.5\height}{\includegraphics[height=3cm]{red-7pt-45679108-LRpath.pdf}} ~~~~~\Rightarrow~~~~~ \lambda_1 \propto \lambda_7\,. \label{fig:nosupport} \end{equation} Since the momenta on legs 1 and 7 are not generically parallel, this cell is not supported by the delta functions. This feature can also be seen from the permutations associated with such cells. Going from twistors to momentum twistors in Section \ref{s:TandMT}, we required that the $B$ plane contains the $\lambda$ 2-plane; see equation \eqref{Cfixed}. If $B$ does not contain a generic 2-plane, this imposes a constraint on the external momenta; in other words, this cell does not intersect the delta function support for generic external momenta. In terms of the permutations, the requirement that $B$ contains a generic 2-plane is that $(\widetilde\sigma(i)-i)\geq 2$ for all $i$ (no column may be in the span of its nearest neighbor, and in particular no column may be zero). To see why, suppose that there exists some $i$ such that $(\widetilde\sigma(i)-i )= 0~\text{or}~1$. Then $b_i \in \text{span}\{b_{i+1}\}$ where $b_i$ is the $i^\text{th}$ column of $B$. Any 2-plane $\lambda\subset B$ would have to satisfy $\<i ,i+1\>=0$, which is clearly not satisfied for generic momenta. Furthermore, recall that the overall Jacobian of the transformation from twistors to momentum twistors, which is just the MHV superamplitude (see equation \eqref{momgrass5}), is singular precisely when $\<i,i+1\>=0$ for some $i$. In the example permutation above, $\widetilde\sigma(7) = 8$, so for the $\lambda$ plane to be contained in this cell it would have to satisfy $\<71\>=0$. Hence such a cell cannot have a momentum twistor dual. Given that the number of cells that meet the support of the delta functions is identical in both spaces, it is not surprising that the permutations which label such cells are related. The permutations for supported momentum space cells $\widetilde\sigma$ can be obtained directly from momentum twistor labels $\sigma$ by the following map: \begin{align}\label{eq:shift} \widetilde{\sigma}(i) = \sigma(i+1) + 1 \,. \end{align} The inverse map from momentum space permutations to momentum twistor permutations, when it exists, was presented in eq.~(8.25) of \cite{ArkaniHamed:2012nw}. Since the momentum space permutations have physically meaningful on-shell diagram representatives, this map also provides a way to associate representative on-shell diagrams with momentum twistor cells of dimension $4$.\footnote{A diagrammatic representation directly in momentum twistor space has also been recently developed \cite{HeBai}.} It is perhaps more surprising that non-vanishing residues in momentum space can also be labeled by $(n-5)$-index sequences similar to those which label the momentum twistor residues \eqref{5res}. From \eqref{minors}, we see that the vanishing of the $i^\text{th}$ minor in momentum twistor space implies that the $(i-1)^\text{th}$ (consecutive) minor in momentum space vanishes as well. Since there are no composite singularities in NMHV momentum twistor space, those residues are uniquely labeled by the set of vanishing minors, $\{i_1,i_2,\ldots,i_{n-5}\}_C$ (not to be confused with a permutation label). By \eqref{minors}, we can label the non-vanishing momentum space residues by a similar list of vanishing momentum space minors: \begin{align} \{i_1,i_2,\ldots,i_{n-5}\}_C \sim \{i_1-1,i_2-1,\ldots,i_{n-5}-1\}_B\,. \end{align} However, setting a collection of $(n-5)$ distinct consecutive minors to vanish in momentum space does not uniquely specify a cell in the Grassmannian. Instead one obtains a union of cells, of which exactly one will have kinematical support. For example, suppose $n = 7$ and we take the cell labeled by $\sigma = \{2,3,4,5,8,6,7\}$, given by the vanishing of the $6^\text{th}$ and $7^\text{th}$ minors. Then by \eqref{eq:shift}, we have $\widetilde{\sigma} = \{4,5,6,9,7,8,10\}$, and the minors labeled by columns $(5,6,7)$ and $(6,7,1)$ vanish. There is exactly one other cell of dimension $2n-4 = 10$ for which exactly these same minors vanish, namely $\widetilde{\sigma}' = \{4,5,6,8,9,7,10\}$. However, $\widetilde{\sigma}'$ does not have kinematical support. This can be seen directly from the two corresponding on-shell diagrams, \begin{equation} \centering \begin{array}{cc} \raisebox{-0.5\height}{\includegraphics[height=3cm]{red-7pt-45697810.pdf}} ~~&~~ \raisebox{-0.5\height}{\includegraphics[height=3cm]{red-7pt-45689710.pdf}}\\ \widetilde{\sigma} = \{4,5,6,9,7,8,10\} ~~~~~&~~~~~ \widetilde{\sigma}' = \{4,5,6,8,9,7,10\} \,. \end{array} \label{fig:nosupport2} \end{equation} The second one vanishes for generic external data since it requires $\lambda_6 \propto \lambda_7$. In general, the momentum space residue $\{i_1-1,i_2-1,\ldots,i_{n-5}-1\}_B$ is just the residue for the cell $\widetilde{\sigma}$. Recalling the Jacobian from Section \ref{s:TandMT}, we have the following relationship between residues in momentum twistor and momentum space: \begin{align} \mathcal{A}_n^{\text{MHV}}\{i_1,i_2,\ldots,i_{n-5}\}_C = \{i_1-1,i_2-1,\ldots,i_{n-5}-1\}_B\,. \end{align} It is also suggestive that, even though the general contours may be difficult to handle, the residues relevant for physics may be easier to study since they will not involve any composite residues. We leave this for future work. \subsubsection{Pushing the boundaries} \label{s:bdr} In Section \ref{s:resapp} we used the boundary operation in two different contexts: one encoded the residue theorems that give linear relations among 5-brackets, the other selected residues of poles in the five brackets. Let us now see how these arise as boundary limits in the momentum twistor space Grassmannian. Let us begin with a 4d cell in the NMHV Grassmannian G$(1,n)$. It is characterized by having precisely five non-vanishing entries in the representative matrix, say $c_{a,b,c,d,x}$. The boundaries of this 4d cell are the 3d cells obtained by setting one extra entry of the representative matrix to zero, i.e.~there are five 3d boundaries. Suppose we go on the boundary characterized by $c_x = 0$. Then there is not generically support on the delta functions in the Grassmannian integral, because they then enforce $c_a Z_a + c_b Z_b + c_c Z_c + c_d Z_d = 0$. This is a constraint on the external data that requires the momentum twistors $Z_{a,b,c,d}$ to be linearly dependent. That is equivalent to the statement that the 4-bracket $\four{abcd}$ vanishes and, as we know from \reef{resresult}, this is precisely one of the poles in 5-bracket $\five{abcdx}$ associated with the residue of our 4d cell. Similarly, we see that the five poles of $\five{abcdx}$ are precisely in 1-1 correspondence with the 3d boundaries of the corresponding 4d cell. This justifies the terminology ``boundary operation" used in the discussion in Section \ref{s:tree}. The relative signs in \reef{bdr5bracket} come from the orientations of the boundaries. Two 4d cells labeled by non-vanishing entries $c_{a,b,c,d,x}$ and $c_{a,b,c,d,y}$, respectively, share one common 3d boundary characterized by $c_x = 0$ and $c_y=0$. As we know from the analysis in Section \ref{s:tree}, the residue of the associated pole at $\four{abcd}$ cancels in the difference of the associated 5-brackets, $\five{abcdx}-\five{abcdy}$. Analogously, the shared boundary between the cells cancels in the sum if they are oppositely oriented. The locality condition of having no pole $\two{abcd}$ for $(a,b,c,d)$ not of the form $(i,i+1,j,j+1)$ translates to the requirement that all shared boundaries not of the corresponding form must be oppositely oriented. These locality conditions, together with the fact that the amplitude is a sum of 4d cells, determine the formula \reef{NMHVtree}. The residue theorems \reef{bdrOp} also have a boundary interpretation in the Grassmannian: boundaries of 5d cells give different equivalent ways of writing the same amplitude formula. We interpret the 4d cells associated with the 5-bracket residues as the boundary of a 5d cell defined by having precisely six non-vanishing entries in the representative matrices, say $c_{a,b,c,d,e,f}$. The six 4d boundaries of this 5d cell are associated with sending one of these six entries to zero. Meanwhile, the Grassmannian integral on the 5d cell is a contour integral on $\mathbb{CP}^1$ with six poles, corresponding to the 4d boundaries above. The sum of the residues at these six poles must therefore be zero by Cauchy's theorem. This is our familiar residue theorem \reef{resCon}, and we see now why it is natural to associate it with a boundary operation. \section{3d ABJM Grassmannian}\label{sec:3DGrass} The fascinating relation between cells of Grassmannian and scattering amplitudes of 4d $\mathcal{N}=4$ SYM has a parallel in 3d ABJM theory \cite{WestCSM,EastCSM}. Previously, an ABJM Grassmannian was developed for external data in momentum space \cite{LeeOG,HW,HWX,KimLee}. The purpose of this section is to apply the strategy from Section \ref{s:TandMT} to derive the ABJM Grassmannian in momentum twistor space. \subsection{Momentum space ABJM Grassmannian} ABJM theory is a 3-dimensional $\mathcal{N}=6$ superconformal Chern-Simons matter theory with 4 complex fermions $\psi^A$ and 4 complex scalars $X^A$, transforming in the $\bf{4}$ and $\bf{\bar{4}}$ under the $SU(4) \sim SO(6)$ R-symmetry. The physical degrees of freedom are the matter fields, and the symmetries imply that the only non-vanishing amplitudes have even multiplicity, in particular one can show that $n=2k+4$ for the N$^k$MHV sector. (For further discussion of 3d kinematics and ABJM amplitudes, see Chapter 11 of the review \cite{Elvang:2013cua}.) A 3d momentum vector $p_i$ can be encoded in a symmetric $2 \times 2$ matrix $p_i^{ab}$. The on-shell condition requires it to have vanishing determinant, so we can write $p_i^{ab} = {\lambda'}_i^a {\lambda'}_i^b$, with $a,b=1,2$ being $SL(2,\mathbb{R})$ indices. One version of the 3d spinor helicity formalism uses these 2-component commuting spinors $\lambda'$ to encode the two on-shell degrees of freedom needed for a null 3d momentum vector. As in 4d, we can form the antisymmetric angle-bracket product $\<ij\> := \epsilon_{ab} {\lambda'}_i^a {\lambda'}_j^b$, although in 3d there are no square-spinors. The $\mathcal{N}=6$ on-shell superspace for 3d ABJM theory involves three Grassmann variables ${\eta'}_{iI}$ with $SU(3) \subset SU(4)$ R-symmetry indices $I=1,2,3$. We are denoting the $\lambda'_i$s (and ${\eta'}_{iI}$s) with primes to distinguish them from a different formulation of the 3d spinor helicity formalism to be introduced in Section \ref{s:MT3d}. In on-shell superspace, the states of ABJM theory are organized in two on-shell supermultiplets \begin{equation} \begin{split} \Phi ~=&~ X_4+\eta'_A\,\psi^A -\frac{1}{2}\epsilon^{ABC}\,\eta'_A\eta'_B\,X_C -\eta'_1\eta'_2\eta'_3\,\psi^4\,,\\ \bar{\Psi} ~=&~\bar{\psi}_4+\eta'_A\bar{X}^A -\frac{1}{2}\epsilon^{ABC}\,\eta'_A\eta'_B\,\bar{\psi}_C -\eta'_1\eta'_2\eta'_3\,\bar{X}^4\,. \end{split} \label{ABJMmap} \end{equation} The superfield $\Phi$ is thus bosonic in nature while $\Psi$ is fermionic. The states of the color-ordered tree-level superamplitude in planar ABJM are arranged with alternating $\Phi$ and $\Psi$, e.g.~$\mathcal{A}_n(\Phi \Psi \Phi \Psi \dots)$, and as a result they do not obey the cyclic invariance $i \rightarrow i+1$ of superamplitudes in planar ${\cal N}=4$ SYM. Instead, the planar ABJM superamplitudes are invariant under $i \rightarrow i+2$, up to a sign of $(-1)^{n/2+1}$. The leading singularities of ABJM theory, and consequently the tree-level amplitudes, enjoy an $OSp(6|4)$ Yangian symmetry \cite{TillDC,ABJMBCFW}, and are given as residues of the following Grassmannian integral \cite{LeeOG,HW,HWX,KimLee}: \begin{equation}\label{MasterInt} \mathcal{L}_{2\tilde{k};\tilde{k}}=\int \frac{d^{2\tilde{k}^2} B'}{GL(\tilde{k})}\,\frac{\delta^{\tilde{k}(\tilde{k}+1)/2}\big(B'\cdot B^{'T}\big) \delta^{2\tilde{k}|3\tilde{k}}\big(B'\cdot \Lambda' \big)}{m'_1 m'_2 \cdots m'_{\tilde{k}} }\,, \end{equation} where $\tilde{k}=k+2 = \frac{n}{2}$ and $\Lambda'_i=(\lambda'_i | \eta'_i)$ is the external data given in 3d momentum space. The denominator contains the product of the first $\tilde{k}$ consecutive minors $m'_i$ of $B'$.\footnote{We note here that for $k$ even this means that the states of the superamplitude are $\mathcal{A}_n(\Phi \Psi \Phi \Psi \dots)$ while for $k$ odd they are $\mathcal{A}_n(\Psi \Phi \Psi \Phi \dots)$. This ensures the correct little group scaling in 3d, for which the superamplitude is invariant for $\Phi$ states and changes signs for $\Psi$ states. Alternatively, one can replace $m_1 m_2 \cdots m_{\tilde{k}}$ by $\sqrt{m_1 m_2 \cdots m_n}$. The two forms are equivalent up to signs depending on the branch of solutions to the orthogonal constraint \cite{HW}. } The bosonic delta-functions enforce the $\tfrac{1}{2}\tilde{k}(\tilde{k}+1)$ constraints \begin{equation} 0= B'\cdot {B'}^{T} = \sum_iB'_{\alpha i}B'_{\beta i} = \sum_{i,j} B'_{\alpha i}B'_{\beta j} g^{ij}\,, \end{equation} with $g^{ij}=\delta^{ij}$ is the trivial metric and $\alpha, \beta = 1,2,\ldots,\tilde{k}$. Thus, in momentum space, the Grassmannian for ABJM theory is an {\bf\em orthogonal Grassmannian} (also known as an {\bf\em isotropic Grassmannian} in the mathematics literature) defined as the space of null $\tilde{k}$-planes in an $n$-dimensional space equipped with an internal metric $g^{ij}$. The metric is trivial in momentum space. We will denote an orthogonal Grassmannian as OG$(\tilde{k},n)$.\footnote{We remark that in the theory of planar electrical networks, the orthogonal Grassmannian defined with respect to the metric $g^{ij} =\delta_{i,j+\tilde k} + \delta_{i,j-\tilde k}$ appears \cite{HS,Lam}. Curiously, the combinatorics of ABJM amplitudes and of planar electrical networks are very closely related, but there is as yet no conceptual explanation of this.} Because of the quadratic condition $B'\cdot {B'}^{T} = 0$, the orthogonal Grassmannian has two distinct components. The dimension of the integral \reef{MasterInt} is \begin{equation} 2 \tilde{k}^2 - \tilde{k}^2 - \frac{1}{2}\tilde{k}(\tilde{k}+1) - 2\tilde{k} +3 = \frac{(\tilde{k}-2)(\tilde{k}-3)}{2} \,, \label{dimInt3dA} \end{equation} where the ``$+3$" is because momentum conservation is automatically encoded in the bosonic delta-functions; this will become evident in the following. Note that in general the metric for a Grassmannian need not be diagonal nor proportional to the identity. For example, positivity for cells of OG$(\tilde{k},n)$ is defined with the metric of alternating signs $(+,-,+,\cdots, -)$ \cite{HWX}. In the following we will see that in converting \reef{MasterInt} into momentum twistor space, we will naturally encounter more general metrics $g$ and denote the corresponding orthogonal Grassmannian as OG$_g(\tilde{k},n)$. \subsection{3d momentum twistors} \label{s:MT3d} We would now like to introduce 3d momentum twistors.\footnote{See~\cite{Lipstein} for an alternative definition.} A natural way to define momentum twistor variables in 3d is to reduce it from 4d. With the 4d conformal group $SO(2,4) \sim SU(4)$, a natural way to introduce momentum twistors is to first define 4d spacetime as a projective plane in 6d. This ``embedding space" formalism~\cite{Embedding} introduces a 6d coordinate $Y^{AB}$, which is anti-symmetric in the $SU(4)$ indices $A,B$. 4d spacetime is then defined to be the subspace $Y^2=\epsilon_{ABCD}Y^{AB}Y^{CD}=0$ with the projective identification $Y\sim rY$ (for $r$ a real or complex number, depending on the context). A solution to the constraint $Y^2=0$ is to write $Y^{AB}$ as a bi-twistor: \begin{equation} Y^{AB}_i=Z_i^{[A}Z_{i-1}^{B]}\,. \end{equation} To honor the projective constraint $Y_i\sim rY_i$, the $Z_i$s must be defined projectively, $Z_i \sim r Z_i$. Here $i=1,2,\dots,n$ label $n$ points $Y_i$ in the embedding space. Consider now the 3d analogue for which the embedding space is 5d. We can start with the 6d space and introduce an $SO(2,3)$-invariant constraint to remove the extra degree of freedom. A natural choice is to impose a $SO(2,3)\sim Sp(4)$ tracelessness condition on $Y$: \begin{equation} Y^{AB}\Omega_{AB}=0\,, \hspace{1cm} \Omega_{AB} = \left(\begin{array}{cc}0 & -I \\ I& 0\end{array}\right)\,. \end{equation} This also implies that the bi-twistors $Z_i$ must satisfy: \begin{equation}\label{ZConstraint} Z_{i}^AZ^B_{i+1}\Omega_{AB}=0\,. \end{equation} Note that \reef{ZConstraint} is projectively well-defined, so we can construct the 3d momentum twistor as a familiar 4d momentum twistor $Z_i=(\lambda_i , \mu_i)$ subject to the constraint: \begin{equation}\label{ZConstraint2} \langle\!\langle i, i{+}1\rangle\!\rangle \,:=\, Z_{i}^AZ^B_{i+1}\Omega_{AB}= \langle \mu_{i} \lambda_{i+1}\rangle-\langle \mu_{i+1}\lambda_{i}\rangle \,=\,0\,. \end{equation} We now need to identify the relation between $\lambda_i$ and $\lambda'_i$. Recall that in three dimensions, a massless momentum can be parameterized as $p_i=E_i(1, \sin\theta_i, \cos\theta_i)$, where $E_i$ is the energy. In bi-spinor notation, we can deduce: \begin{equation} \lambda'_{ia}\lambda'_{ib}=p_{iab}=E_i\left(\begin{array}{cc}-1+\cos\theta_i & \sin\theta_i \\ \sin\theta_i & -1-\cos\theta_i\end{array}\right)\quad \rightarrow \quad \lambda'_{ia}=i\sqrt{2E_i}\left(\begin{array}{c} -\sin\frac{\theta_i}{2} \\ \cos\frac{\theta_i}{2} \end{array}\right)\,. \end{equation} Now since $Z_i$ is defined projectively, the components of $Z_i$ must have well-defined projective scalings. This is not possible with $Z_i=(\lambda'_i,\mu_i)$ because $p_i$ is not invariant under the scaling $\lambda'_i \rightarrow t_i \lambda'_i$. Consequently, $y_i=p_{i}-p_{i+1}$ cannot have any nice homogenous scaling property and neither can $\mu_i$, since the latter is defined through the incidence relation $\mu_i^a=y_i^{ab}\lambda'_{i b}$. The resolution is to parameterize the 3d kinematics in a fashion that is similar to 4d. We define \begin{equation}\label{NewVariable} \lambda_{i a}=\left(\begin{array}{c} -\sin\frac{\theta_i}{2} \\ \cos\frac{\theta_i}{2} \end{array}\right)\,, \hspace{1cm} \tilde{\lambda}_{i a}=-2E_i\lambda_{i a} \end{equation} such that we now have $p_i=\lambda_i\tilde{\lambda}_i$. For simplicity, we set $\tilde{E}_i=-2E_i$ in the following. Note that the number of degrees of freedom for each particle is still 2 and that $p_i$ is now invariant under the following scaling rules: \begin{equation} \lambda_i\rightarrow t_i\lambda_i, \quad \tilde{E}_i\rightarrow t_i^{-2} \tilde{E}_i\,. \end{equation} Since $p_i$ (and hence $y_i$) is invariant, $\mu_i$ has the same scaling property as $\lambda_i$ through the new incidence relation \begin{equation}\label{Incidence2} \mu_i^a= y_i^{ab}\lambda_{i b} =y_{i+1}^{ab}\lambda_{i b}: ~~\quad \mu_i\rightarrow t_i\mu_i\,. \end{equation} With the incidence relation (\ref{Incidence2}) and the symmetry of the $y_i$-matrices, the constraint (\ref{ZConstraint2}) is automatically satisfied \begin{equation} \langle \mu_{i} \lambda_{i+1}\rangle-\langle \mu_{i+1}\lambda_{i}\rangle =\lambda_{ia}y_i^{ab}\lambda_{i{+}1,b} -\lambda_{i+1,a}y_{i{+}1}^{ab}\lambda_{ib}=0\,. \end{equation} Thus we have deduced a suitable form of 3d momentum twistor which has a well defined projective property. It has many of the same properties as its 4d cousin, for example \reef{y-lambda-mu} holds, and one can directly derive the 3d versions of the relations \reef{inc}: \begin{equation} \begin{split} \tilde{\lambda}_{i} &=~ \frac{\two{i+1,i}\mu_{i-1} + \two{i,i-1}\mu_{i+1} + \two{i-1,i+1}\mu_{i} } {\two{i-1,i}\two{i,i+1}}\,, \\[2mm] \tilde{\eta}_{iA} &=~ \frac{\two{i+1,i}\eta_{i-1,A} + \two{i,i-1}\eta_{i+1,A} + \two{i-1,i+1}\eta_{iA} } {\two{i-1,i}\two{i,i+1}}\,. \end{split} \label{inc3d} \end{equation} where we define $\tilde{\eta}_i:=\eta_i'\sqrt{\tilde{E}_i}$. In summary, we have found that the new momentum space variables \begin{equation} \lambda_i=\frac{1}{\sqrt{\tilde{E}_i}}\lambda'_i\,, ~~~~~~~~ \tilde{\lambda}_i=\sqrt{\tilde{E}_i}\lambda'_i\, , ~~~~~~~~ \tilde{\eta}_i=\eta_i'\sqrt{\tilde{E}_i} \end{equation} facilitate the introduction of 3d momentum supertwistors $\mathcal{Z}_i = (\lambda_i, \mu_i | \eta_i)$, which are just like the 4d ones but subject to the constraints \reef{ZConstraint2}. Also, in addition to the $SL(4)$-invariant $\<ijkl\>$ defined in \reef{def4bracket}, we now have a 2-bracket invariant \begin{equation} \langle\!\langle ij\rangle\!\rangle:= Z_{i}^AZ^B_{j}\Omega_{AB}\, . \label{angle2} \end{equation} As noted in \reef{ZConstraint2}, the projection from 4d momentum twistors to 3d ones is defined by $\langle\!\langle i, i+1\rangle\!\rangle =0$. The 4-brackets and 2-brackets are related via a version of the Schouten identity: \begin{equation} \langle ijkl\rangle=\langle\!\langle ij\rangle\!\rangle\langle\!\langle kl\rangle\!\rangle+\langle\!\langle ik\rangle\!\rangle\langle\!\langle lj\rangle\!\rangle+\langle\!\langle il\rangle\!\rangle\langle\!\langle jk\rangle\!\rangle \,. \label{4Dschouten} \end{equation} This follows from the identity $\epsilon_{ABCD}=-\Omega_{AB}\Omega_{CD}-\Omega_{AC}\Omega_{DB}-\Omega_{AD}\Omega_{BC}$. Note that the RHS of \reef{4Dschouten} has the same form as the 2d Schouten identity for angle-brackets, but the LHS has the non-vanishing contraction with the Levi-Civita symbol because the momentum twistors are 4-component objects. As a side-remark, we note that just as the parameterization in (\ref{NewVariable}) can be viewed as a descendant from the 4d spinor helicity variables, it also has another 2d sibling: the 2d momentum twistors that correspond to taking $\theta=0$ or $\pi$ in (\ref{NewVariable}): \begin{equation} \lambda_i|_{\theta_i=0}=\left(\begin{array}{c} 0 \\ 1 \end{array}\right)=: \lambda^+, ~~~~ \tilde\lambda^+= \tilde{E}_i\lambda^+, ~~~~ \lambda_i|_{\theta_i=\pi}=\left(\begin{array}{c}-1 \\ 0\end{array}\right)=: \lambda^-, ~~~~ \tilde\lambda^-= \tilde{E}_i\lambda^-\,. \end{equation} The superscript $(+,-)$ indicates which of the two distinct light-cone directions contains the corresponding momentum: \begin{equation} \lambda_i^+\tilde{\lambda}_i^+ =p_i^+ =\left(\begin{array}{cc}0 & 0 \\0 & \tilde{E}_i\end{array}\right), \quad ~~~ \lambda_i^-\tilde{\lambda}_i^- =p_i^- =\left(\begin{array}{cc}\tilde{E}_i & 0 \\0 & 0\end{array}\right)\,. \end{equation} These 2d momentum variables have been used to study scattering amplitudes of planar $\mathcal{N}=4$ SYM limited to 2d kinematics~\cite{2DK1, 2DK2}. \subsection{Derivation of ABJM momentum twistor space Grassmannian} We are now ready to convert the integral formula in (\ref{MasterInt}) to momentum twistor space.\footnote{A similar attempt was initiated in~\cite{UnPub} with a different definition of 3d momentum twistors. As a result, projectivity was not well defined.} First we change variables $B'_{\alpha i} = \sqrt{\tilde{E}_i} B_{\alpha i}$, $\lambda'_i =\tilde{\lambda}_i / \sqrt{\tilde{E}_i}$, $\eta_i'=\tilde{\eta}_i/\sqrt{\tilde{E}_i}$, so that only the variables in (\ref{NewVariable}) appear in the bosonic delta functions: \begin{equation} \mathcal{L}_{2\tilde{k};\tilde{k}}= J_E \int \frac{d^{2\tilde{k}^2}B}{GL(\tilde{k})\,m_1 m_2 \cdots m_{\tilde{k}} } \,\delta^{\tilde{k}(\tilde{k}+1)/2}\big(B_{\alpha i}B_{\beta j}g^{ij}\big) \,\delta^{2\tilde{k}}\big(B\cdot \tilde\lambda\big) \,\delta^{(3\tilde{k})}\big(B\cdot \tilde{\eta}\big)\,, \end{equation} where the factor \begin{equation} J_E = \frac{\prod_{i=1}^{n} \tilde{E}_i^{\tilde{k}/2}} {\big( \tilde{E}_1 \tilde{E}_2^2 \tilde{E}_3^3 \cdots \tilde{E}_{\tilde{k}}^{\tilde{k}} \tilde{E}_{\tilde{k}+1}^{\tilde{k}-1} \cdots \tilde{E}_{n-2}^2 \tilde{E}_{n-1} \big)^{1/2} } \end{equation} comes from the scaling of the measure and the minors. The $2\tilde{k}$-dimensional metric $g^{ij}$ is \begin{equation} g^{ij}=\left(\begin{array}{cccc}\tilde{E}_1 & 0 & 0 & 0 \\0 & \tilde{E}_2 & 0 & 0 \\0 & 0 & \ddots & 0 \\0 & 0 & 0 & \tilde{E}_{2k}\end{array}\right)\,. \end{equation} Just as in 4d, the delta function $\delta^{2\tilde{k}}(B\cdot \tilde\lambda)$ requires $B$ to be orthogonal to the $\tilde{\lambda}$-plane, and by momentum conservation ($\sum_i \lambda_i \tilde\lambda_i = 0$), the $B$-plane must therefore contain the $\lambda$-plane. We can use this to gauge-fix part of the $GL(\tilde{k})$ as in 4d: \begin{equation} B_{\alpha i}= \left(\begin{array}{c} B_{\hat\alpha i} \\ \lambda^1_i \\ \lambda^2_i \end{array}\right)\,, \end{equation} where $\hat\alpha=1,\ldots,k$ and $k=\tilde{k}-2$. With this gauge choice, the remaining delta functions become\footnote{Note that for $\alpha=\tilde{k}-1,\tilde{k}$, we have $\lambda_i B_{\hat{\beta}j}g^{ij}=\lambda_i B_{\hat{\beta}i}\tilde{E}_i=\tilde{\lambda}_i B_{\hat{\beta}i}$} \begin{eqnarray} \delta^{\tilde{k}(\tilde{k}+1)/2}\big(B_{\alpha i}B_{\beta j}g^{ij}\big) \,\delta^{3\tilde{k}}\big(B\cdot\tilde{\eta}\big) \rightarrow \delta^{3}(\mathcal{P})\delta^{6}(\mathcal{Q}) \,\delta^{k(k+1)/2}\big(B_{\hat{\alpha}i}B_{\hat\beta j}g^{ij}\big) \,\delta^{2k|3k}\big(B_{\hat{a}}\cdot\tilde\Lambda\big)\,, \end{eqnarray} where $\tilde\Lambda=(\tilde\lambda | \tilde{\eta})$, $\mathcal{P}$ is the total momentum, and $\mathcal{Q}$ is the total supermomentum. Now we can follow the steps from the $\mathcal{N}=4$ SYM analysis in Section \ref{s:TandMT} to convert the Grassmannian integral to one with momentum twistor external data: \begin{enumerate} \item The relation between the momentum space data and the momentum twistor variables allow us to introduce a new variable $C_{\hat{\alpha} i}= B_{\hat\alpha j}Q_{ji}$, with $Q$ defined in \reef{CfromBgauge12} (see also \reef{Qmatrix}). \item Rewrite the minors using \reef{minors}. \item To invert the relation $B_{\hat\alpha j}Q_{ji}$, we use translation invariance $T_k$ to fix $B_{\hat{\alpha} 1}=B_{\hat{\alpha} 2}=0$, thus obtaining $B_{\hat\alpha \hat{j}}=C_{\hat\alpha \hat{i}}(Q^{-1})_{\hat{i}\hat{j}}$, with $\hat{i}=3,\ldots, n$. A simple expression for $Q^{-1}$ is given in \eqref{Qinv}, and we verify its form in Appendix \ref{app:Qinv}. This allows us to change variables from $B_{\hat\alpha \hat{j}}$ to $C_{\hat\alpha \hat{j}}$. We restore the integration variables to include $C_{\hat\alpha 1}, C_{\hat\alpha 2}$ by introducing $\delta^{2k}(C\cdot \lambda)$. \end{enumerate} This brings us to the following final form of the Grassmannian integral ($n=2k+4$): \begin{equation}\label{MasterInt2} \mathcal{L}_{n;k} = J \times \delta^{3}(\mathcal{P})\,\delta^{(6)}(\mathcal{Q}) \times \int \frac{d^{kn}C}{GL(k)} \frac{\delta^{k(k+1)/2}\Big(C_{\hat\alpha \hat{i}}(Q^{-1})_{\hat{i}\hat{j}}C_{\hat\beta\hat{k}}(Q^{-1})_{\hat{k}\hat{l}}\,g^{\hat{j}\hat{l}}\Big) \,\delta^{4k|3k}\big(C\cdot \mathcal{Z}\big)}{M_2 M_3 \cdots M_{k+3}}\,, \end{equation} where $\mathcal{Z}_i=(Z_i|\eta_i)$ and \begin{equation} J = \big(\langle 12\rangle\langle 23\rangle\cdots \langle n1\rangle\big)^k \frac{ \prod_{i=1}^{n} \tilde{E}_i^{(k+2)/2}} { \tilde{E}_{k+2}^{(k+2)/2} \prod_{i=1}^{k+1} \Big(\tilde{E}_i^{1/2} \<i,i+1\> \<n-i-1,n-i\>\tilde{E}_{n-i}^{1/2} \Big)^{i}}\,. \label{uglyJ} \end{equation} It is straightforward to verify that the $GL(1)$ weight of the integral in \reef{MasterInt2} cancels. It should be noted that here and in the remainder of this section, the angle-brackets $\<ij\>$ are now composed of the $\lambda$-spinors, not the $\lambda'$'s. There is an unsatisfactory feature in (\ref{MasterInt2}): the sum in the constraint $C Q^{-1} C Q^{-1} g=0$ only runs over $\hat{i},\hat{j}=3,\ldots, n$, and therefore $(Q^{-1} Q^{-1} g)^{\hat{i}\hat{j}}$ cannot be interpreted as an orthogonal constraint on the Grassmannian $C$. This is because an orthogonal Grassmannian is defined with a metric $G^{ij}$ whose indices run over the full $n$-dimensional space. However, using the delta function support we can rewrite $C Q^{-1} C Q^{-1} g$ in terms of a non-degenerate effective metric $G^{ij}$ specified by the external data. To see this note that on the support of the $\delta(C\cdot \lambda)$, we have \begin{equation}\label{DQ} C_{\hat{\alpha} \hat{i}}(Q^{-1})_{\hat{i}\hat{j}}=\sum_{i=2}^{\hat{j}-1}C_{\hat{\alpha} i}\langle i\hat{j}\rangle\,. \end{equation} For example, for $k=1$, we have $\hat{i},\hat{j}=3,\ldots,6$ and \begin{equation} C_{\hat{\alpha} \hat{i}}(Q^{-1})_{\hat{i}\hat{j}}= \left(\begin{array}{c} C_{\hat{\alpha} 2}\langle 23\rangle \\ C_{\hat{\alpha} 2}\langle 24\rangle +C_{\hat{\alpha} 3}\langle 34\rangle\\ C_{\hat{\alpha} 2}\langle 25\rangle +C_{\hat{\alpha} 3}\langle 35\rangle+C_{\hat{\alpha} 4}\langle 45\rangle \\ C_{\hat{\alpha} 2}\langle 26\rangle +C_{\hat{\alpha} 3}\langle 36\rangle+C_{\hat{\alpha} 4}\langle 46\rangle+C_{\hat{\alpha} 5}\langle 56\rangle \end{array}\right)\,. \label{CQ1} \end{equation} Note that $C_{\hat{\alpha} 1}$ and $C_{\hat{\alpha} 6}$ do not appear in (\ref{CQ1}). We can use $C\cdot \lambda=0$ to get an expression in terms of a `conjugate' set of $C_{\hat{\alpha} i}$'s, e.g.~\begin{equation} C_{\hat{\alpha} \hat{i}}(Q^{-1})_{\hat{i}\hat{j}} = -\left(\begin{array}{c} C_{\hat{\alpha} 1}\langle 13\rangle +C_{\hat{\alpha} 4}\langle 43\rangle+C_{\hat{\alpha} 5}\langle 53\rangle+C_{\hat{\alpha} 6}\langle 63\rangle \\ C_{\hat{\alpha} 1}\langle 14\rangle +C_{\hat{\alpha} 5}\langle 54\rangle+C_{\hat{\alpha} 6}\langle 64\rangle \\ C_{\hat{\alpha} 1}\langle 15\rangle +C_{\hat{\alpha} 6}\langle 65\rangle \\ C_{\hat{\alpha} 1}\langle 16\rangle\end{array}\right)\,. \label{CQ2} \end{equation} To reveal the symmetric form of the effective (inverse) metric $\tilde{G}^{ij}$ defined via $C_{\hat{\alpha} i} C_{\hat{\beta} j} \tilde{G}^{ij} = ( C_{\hat{\alpha}} Q^{-1}) ( C_{\hat{\alpha}} Q^{-1})g$, we take the symmetric (in $\hat{\alpha}$ and $\hat{\beta}$) product of two copies of $C_{\hat{\alpha} \hat{i}}(Q^{-1})_{\hat{i}\hat{j}}$, one in the form \reef{CQ1} and one in the conjugate form \reef{CQ2}. For $k=1$, we find \begin{equation} \tilde{G}^{ij} = \frac{1}{2} \left(\begin{array}{cccccc}0 & 0 & -\langle 1|2|3\rangle & -\langle 1|2+3|4\rangle & -\langle 1|2+3+4|5\rangle & 0 \\ 0 & 0 & 0 & \langle 2|3|4\rangle & \langle 2|3+4|5\rangle & \langle 2|3+4+5|6\rangle \\ * & 0 & 0 & 0 & \langle 3|4|5\rangle & \langle 3|4+5|6\rangle \\ * & * & 0 & 0 & 0 & \langle 4|5|6\rangle \\ * & * & * & 0 & 0 & 0 \\ 0 & * & * & * & 0 & 0\end{array}\right)\,, \end{equation} where the terms denoted by $*$ are related to the ones explicitly written via symmetry, $\tilde{G}^{ji} = \tilde{G}^{ij}$. The notation $\langle i| l |j\rangle=\langle i|p_l|j\rangle$ uses $\lambda_i g^{il} \lambda_l=p_i$ (only $l$ summed over). To rewrite the entries of the effective matrix in terms of the momentum twistors, note that \begin{equation} \langle i| (p_{i}+p_{i+1}+\ldots+p_{j-1} )|j\rangle=\langle i| (y_i-y_j) |j\rangle= Z_{i}^AZ^B_{j}\Omega_{AB}\, =: \, \langle\!\langle ij\rangle\!\rangle \label{doublebracket} \end{equation} The constraints \reef{ZConstraint2} on the external data is $\langle\!\langle i, i+1\rangle\!\rangle=0$, so in terms of the double-bracket \reef{doublebracket}, we can write the effective metric as \begin{equation}\label{GDef} \tilde{G}^{ij}=\langle\!\langle ij\rangle\!\rangle\quad {\rm for}\quad 2 \le i<j \le n \,, \hspace{1cm} \tilde{G}^{1j}=-\langle\!\langle 1j\rangle\!\rangle\quad {\rm for}\quad 2 \le j \le n \,. \end{equation} This defines a non-degenerate metric only for $n>4$, since for $n=4$ the only non-trivial elements are $C_1 \langle\!\langle 13\rangle\!\rangle C_3$ and $C_2\langle\!\langle 24\rangle\!\rangle C_4$ which vanishes under the support of $\delta(C\cdot Z)$. While the metric is non-degenerate for $n>4$, it is not manifestly cyclic symmetric. Let us first inspect the orthogonality condition: \begin{equation} 0 = \sum_{1 \le i , j \le n} C_{\hat{\alpha} i}C_{\hat{\beta} j} \tilde{G}^{ij} = \sum_{3 \le i < j \le n} C_{\hat{\alpha} i}C_{\hat{\beta} j} \langle\!\langle \ij\rangle\!\rangle + (\hat{\alpha} \longleftrightarrow \hat{\beta} ) \,. \end{equation} The second equality is obtained using $C\cdot Z = 0$. Now, to see that the orthogonality constraint is indeed cyclic invariant, one uses $C\cdot Z = 0$ to show that $\sum_{4 \le i < j \le n+1} C_{\hat{\alpha} i}C_{\hat{\beta} j} \langle\!\langle \ij\rangle\!\rangle= \sum_{3 \le i < j \le n} C_{\hat{\alpha} i}C_{\hat{\beta} j} \langle\!\langle \ij\rangle\!\rangle$, with the understanding that $n + 1$ equals $1$. This suffices to prove that the condition $CG C^T = 0$ is cyclic invariant. Next, we can simplify the sum of $n$ cyclic copies of the orthogonality condition to find an equivalent, manifestly cyclic invariant, form of the metric: \begin{equation}\label{GDef1} \left\{\begin{array}{c} G^{i,i{+}2}=\frac{k}{n}\langle\!\langle i,i{+}2\rangle\!\rangle \\ G^{i,i+3}=\frac{k{-}1}{n}\langle\!\langle i,i{+}3\rangle\!\rangle \\ \vdots \\ G^{i,i{+}k{+}1}=\frac{1}{n}\langle\!\langle i,i{+}k{+}1\rangle\!\rangle\end{array}\right. \end{equation} while $G^{ij}=0$ for all other cases. We then have the final form for the cyclically invariant momentum twistor Grassmannian integral for 3d ABJM. It is an orthogonal Grassmannian whose metric \reef{GDef1} depends on the external data: \begin{equation} \label{MasterInt3} \boxed{ \mathcal{L}_{n;k} = J \times \delta^{3}(\mathcal{P})\,\delta^{(6)}(\mathcal{Q}) \times \int \frac{d^{kn}C}{GL(k)}\frac{\delta^{\frac{k(k+1)}{2}}\left( C_{\hat\alpha i} G^{ij}C_{\hat\beta j} \right)\delta^{4k|3k}(C\cdot \mathcal{Z})}{M_2 M_3 \cdots M_{k+3}}\,,} \end{equation} with $n=2k+4$ and the Jacobian $J$ given in \reef{uglyJ}. Note that the integral is indeed projectively invariant under rescaling of $Z_i\rightarrow t_i Z_i$ due to the form of the effective metric $G^{ij}$ in (\ref{GDef}). The momentum twistor space Grassmannian integral for ABJM theory and $\mathcal{N}=4$ SYM both have $4k$ bosonic delta functions $\delta^{4k}(C \cdot Z)$, but in addition the ABJM integral \reef{MasterInt3} has the extra orthogonal constraint. In $\mathcal{N}=4$ SYM, the $\big(k(n{-}k){-}4k\big)$ remaining degrees of freedom in the momentum twistor Grassmannian are localized by the minors. For ABJM, $k(k{+}1)/2$ of the $\big(k(n{-}k){-}4k\big)$ degrees of freedom are localized by the orthogonal constraint, so the dimension of the integral \reef{MasterInt3} is \begin{equation} 2(k+2)k-k^2-4k -\frac{1}{2}k(k+1)=\frac{k(k-1)}{2}\,, \label{dimInt3dB} \end{equation} the same as the dimension \reef{dimInt3dA} of the momentum space Grassmannian integral \reef{MasterInt}. In particular, we note that for $n=6$ (i.e.~$k=1$) the integral localizes completely. Because the orthogonality constraint is quadratic, there are two solution branches that the integral localizes on and we must add them to obtain the $n=6$ tree-level ABJM superamplitude. This matches the observation that there is only one BCFW-diagram for the $n=6$ ABJM amplitude, but the kinematic constraint is quadratic, so the diagram yields a two-term contribution. Those are the two terms given by the two branches of the orthogonal Grassmannian. In the following, we study the orthogonality condition and evaluate the integral \reef{MasterInt3} explicitly for $n=6$. \subsection{The 6-point ABJM amplitude in momentum twistor space} For $n=6$, the integral \reef{MasterInt3} becomes \begin{equation} \mathcal{L}_{6;1} = J_{234} \times \delta^{3}(\mathcal{P})\,\delta^{(6)}(\mathcal{Q}) \int \frac{d^6 c}{GL(1)} \frac{\delta\big( c_i G^{ij} c_j \big)\delta^4 \big( c\cdot Z \big)\delta^{(3)} \big( c\cdot \eta \big)} {c_2 c_{3} c_{4}} \,, \label{L61} \end{equation} where the Jacobian is given by \reef{uglyJ} and is \begin{equation} J_{234} = \frac{\<12\> \<23\>\<34\>\<45\>\<56\> \<61\> \prod_{i=1}^6 \tilde{E}_i^{3/2}} {\tilde{E}_3^{3/2} \tilde{E}_1^{1/2} \<12\> \<45\> \tilde{E}_5^{1/2} \tilde{E}_2 \<23\>^2 \<34\>^2 \tilde{E}_4} \,. \label{J234} \end{equation} If we had picked a representation of the original momentum space Grassmannian integral with a different product of $3$ consecutive minors in the denominator, the integral $\mathcal{L}_{6;1}$ would have a denominator $c_i c_{i+1} c_{i+2}$ and the associated Jacobian $J_{i,i+1,i+2}$ would be obtained from \reef{J234} by relabeling of the lines. \subsubsection{Orthogonality constraint and symmetry under $i \rightarrow i+2$} \label{s:orthog} In the momentum space Grassmannian \reef{MasterInt}, the orthogonality condition $B'\cdot B'^T=0$ implies the following relation among the minors: \begin{equation} m_{i}' m_{i+1}' = (-1)^{\tilde{k}-1} m'_{i+\tilde{k}} m'_{i+1+\tilde{k}} \,, \label{minorID} \end{equation} with $\tilde{k}=k+2$ and indices mod $n$. The relation \reef{minorID} is key for proving that the Grassmannian integral \reef{L61} has the appropriate cyclic invariance under $i \rightarrow i+2$. The equivalent relation for the minors of the momentum twistor space Grassmannian will depend on the external data. Let us work out what it is for $n=6$ and how it can be used to prove that our momentum twistor Grassmannian \reef{L61} has cyclic symmetry $i \rightarrow i+2$. Using $C \cdot Z=0$, direct evaluation of the orthogonality condition gives \begin{eqnarray} \nonumber 0 &=& c_i G^{ij} c_j ~=~ \frac{1}{6}\Big( c_1 \langle\!\langle 13\rangle\!\rangle c_3+c_2\langle\!\langle 24\rangle\!\rangle c_4+c_3\langle\!\langle 35\rangle\!\rangle c_5+c_4\langle\!\langle 46\rangle\!\rangle c_6+c_5\langle\!\langle 51\rangle\!\rangle c_1 +c_6\langle\!\langle 62\rangle\!\rangle c_2 \Big) \\ &=& \frac{1}{2} c_3 \langle\!\langle 35 \rangle\!\rangle c_5 - \frac{1}{2} c_2 \langle\!\langle 26 \rangle\!\rangle c_6\,. \label{n6cGc} \end{eqnarray} Since the metric $G^{ij}$ is cyclic invariant, other forms of the constraint can be obtained from cyclic symmetry: there are three distinct ones: \begin{equation} c_3 \langle\!\langle 35 \rangle\!\rangle c_5 = c_2 \langle\!\langle 26 \rangle\!\rangle c_6\,,~~~~ c_4 \langle\!\langle 46 \rangle\!\rangle c_6 = c_3 \langle\!\langle 31 \rangle\!\rangle c_1\,,~~~~ c_5 \langle\!\langle 51 \rangle\!\rangle c_1 = c_4 \langle\!\langle 42 \rangle\!\rangle c_2\,. \label{n6orthocycl} \end{equation} It follows from the first two identities in \reef{n6orthocycl} that \begin{equation} c_1 c_2 = \frac{\langle\!\langle 35 \rangle\!\rangle \langle\!\langle 46 \rangle\!\rangle}{\langle\!\langle 26 \rangle\!\rangle \langle\!\langle 31 \rangle\!\rangle} c_4 c_5 = \frac{\four{3456}}{\four{6123}} c_4 c_5\,. \label{c1c2TOc4c5} \end{equation} The second equality follows from \reef{4Dschouten}. The property \reef{c1c2TOc4c5} is the equivalent of \reef{minorID} in momentum twistor space. The other relations $c_{i} c_{i+1} \propto c_{i+3} c_{i+4}$ (indices mod 6) are obtained from cyclic relabeling of \reef{c1c2TOc4c5}. Let us now examine the cyclic symmetry $i \rightarrow i+2$. The only part that changes in the integral of \reef{L61} is the product $c_2 c_3 c_4$, which becomes $c_4 c_5 c_6$. It follows from a cyclic version of \reef{c1c2TOc4c5} that \begin{equation} \frac{1}{c_2 c_3 c_4} = \frac{\four{1234}}{\four{4561}} \frac{1}{c_4 c_5 c_6} \end{equation} It is not hard to see that the factor $\four{1234}/\four{4561}$ is exactly compensated by the non-trivial Jacobian: by \reef{J234} and its version with $i \rightarrow i+2$ \begin{equation} \frac{J_{456}}{J_{234}} = \frac{\<12\> \tilde{E}_2 \<23\>^2 \tilde{E}_3 \<34\>}{\<45\> \tilde{E}_5 \<56\>^2 \tilde{E}_6 \<61\>} = \frac{\dtwo{13}\dtwo{24}}{\dtwo{46}\dtwo{51}} =\frac{\four{1234}}{\four{4561}}\,. \end{equation} Here we use the form \reef{doublebracket} of the $Sp(4)$-product to write $\dtwo{i,i+2} = \<i| p_{i+1} | i+2\> = \<i,i+1\> \tilde{E}_{i+1} \<i+1,i+2\>$ and the identity \reef{4Dschouten}. Thus, we conclude that \begin{equation} \frac{J_{234}}{c_2 c_3 c_4} = \frac{J_{456}}{c_4 c_5 c_6} \end{equation} and that the $n=6$ integral is invariant under $i \rightarrow i+2$; the non-trivial orthogonality condition and the overall Jacobian factor nicely conspire to give this result. \subsubsection{Evaluation of the $n=6$ integral} To evaluate the integral \reef{L61}, we first use the bosonic delta function constraints for the six-point momentum twistor space Grassmannian, which fixes four of the integrations. As in Section \ref{s:NMHVeval}, we use the 5-term Schouten identity \reef{5Schouten} to expand the constraint $C\cdot Z=0$ on a basis of four $Z$'s, although now we find it convenient to solve for $c_2,c_3,c_5,c_6$: \begin{equation}\label{cSol} \begin{split} &c_2=\frac{c_1\langle 3561\rangle+c_4\langle 3564\rangle}{\langle 2356\rangle}\,,\; ~~~~ c_3=\frac{c_1\langle 5612\rangle+c_4\langle 5642\rangle}{\langle 2356\rangle}\,, \\[1mm] & c_5=\frac{c_1\langle 6123\rangle+c_4\langle 6423\rangle}{\langle 2356\rangle}\,,\; ~~~~ c_6=\frac{c_1\langle 1235\rangle+c_4\langle 4235\rangle}{\langle 2356\rangle}\,. \end{split} \end{equation} Evaluating the delta-function this way will generate a Jacobian factor of $1/\four{2356}$. The orthogonal constraint \reef{n6cGc} becomes \begin{equation} 0=~ c_i G^{ij} c_j = \frac{\dtwo{42}\four{3456}}{2\four{2356}} \bigg( c_4^2 - c_1^2\frac{\dtwo{31}^2\four{5612}}{\four{1234}\four{3456}} \bigg)\,. \label{n6ortho} \end{equation} We fix the $GL(1)$ gauge by setting $c_1=\four{2356}$; the result is going to be independent of the gauge, but this choice will simplify the other $c_i$'s. The solution to \reef{n6ortho} is then: \begin{equation} c_4^\pm = \pm \frac{\dtwo{31}\four{5612}\four{2356}}{\sqrt{D}}\,, \label{c4pm} \end{equation} where $D= \<1234\>\<3456\>\<5612\>= -\prod_{i=1}^6\langle\!\langle i,i+2\rangle\!\rangle$. The Grassmannian integral localizes on the two solutions \reef{c4pm}: \begin{equation} I_i = \int \frac{\delta\big( c_i G^{ij} c_j \big)\delta^4 \big( c\cdot Z \big)\delta^{(3)} \big( c\cdot \eta \big)} {c_i c_{i+1} c_{i+2}} = \sum_{s=\pm1} \frac{2\four{2356}}{2c_4^s\,\dtwo{42}\four{3456}} \frac{\four{2356}}{\four{2356}} \frac{ \delta^{(3)}\big( c\cdot\eta \big) }{{c_i c_{i+1} c_{i+2}}}\bigg|_{c_4 = c_4^s}\,, \label{Ii0} \end{equation} where $i$ depends on the organization of the external states (i.e.~which 3 minors we selected at the starting point in momentum space). It is implicitly understood that on the RHS, $c_1=\four{2356}$ and $c_{2,3,5,6}$ are given by \reef{cSol} with $c_4$ as in \reef{c4pm}. The factor ${2\<2356\>}/{2c_4^s\,\dtwo{42}\four{3456}}$ is the Jacobian of the orthogonality condition (see footnote \ref{footie:delta}), the second $\four{2356}$ in the numerator is the gauge-fixing Jacobian, and the $1/\<2356\>$ comes from evaluating $\delta^4(c\cdot Z)$. The prefactors readily simplify and we get \begin{equation} I_i ~=~ \frac{\delta^{(3)}\big( c\cdot\eta \big) }{{\sqrt{D}\, c_i c_{i+1} c_{i+2}}}\bigg|_{c_4 = c_4^+} - \frac{\delta^{(3)}\big( c\cdot\eta \big) }{{\sqrt{D}\, c_i c_{i+1} c_{i+2}}}\bigg|_{c_4 = c_4^-}\,. \label{n6integLoc} \end{equation} It is clear from \reef{cSol} and \reef{c4pm} that the individual $c_i$'s are expressions with square roots. It may be worrisome to see such denominator terms arise from the Grassmannian integral; after all, we would expect the denominators to be a product of physical poles. However, the expectation is warranted; for each $i$, just write \begin{equation} \frac{1}{c_i^\pm} = \frac{c_i^\mp}{c_i^\pm c_i^\mp } \,, ~~~~\text{where}~~~~ c_i^\pm := c_i|_{c_4 = c_4^\pm} ~~\text{for}~~ i=2,3,5,6\,. \end{equation} The combinations $c_i^+ c_i^-$ are manifestly free of any square roots. Furthermore, after applications of the Schouten identities \reef{5Schouten} and \reef{4Dschouten}, one finds that each $c_i^+ c_i^-$ has a nice factorized form involving only 4-brackets: \begin{equation} \begin{array}{llll} c_1^+ c_1^- = \<2356\>^2\,,& c_2^+ c_2^- = \frac{\< 5613 \> \< 6134\>\<2356\>}{\<6234\>} \,,& c_3^+ c_3^- = \frac{\< 1245\> \< 5612\>\<2356\>}{\<2345\>}\,,\\[4mm] c_4^+ c_4^- = \frac{\<1235\> \< 5612\>\<2356\>^2}{ \<2456\> \<2345\>} \,,~~~& c_5^+ c_5^- = \frac{\<6123\> \< 6134\>\<2356\>}{\<3456\>} \,,~~~& c_6^+ c_6^- = \frac{\<1235\>\< 1245\> \<2356\>}{\<2456\>}\,. \end{array} \label{cici} \end{equation} Choosing $i=2$ in \reef{n6integLoc} and using \reef{cici}, we find{\footnotesize \begin{align} I_2 = { \sum_{s=\pm}} \frac{\dtwo{24}^2 { \delta^{(3)}\big(c^{(s)}\cdot\eta \big)} \big[ \dtwo{35} \big(\tfour{1236}\tfour{2456}-\tfour{5612}\tfour{2346}\big) + s\sqrt{D}\big(\dtwo{26}\dtwo{35}+\dtwo{25}\dtwo{36}\big) \big] }{\tfour{1234}\tfour{5612}^2\tfour{1245}\tfour{3461}\tfour{2356}^3}.\!\! \label{Is} \end{align} The result \reef{Is} for the 6-point amplitude has two terms because the orthogonal Grassmannian has two branches. To compare, the BCFW calculation of the 6-point amplitude involves only one diagram, but it gives rise to two terms, just as in \reef{Is}, because the on-shell condition for 3d BCFW is not linear in the shift parameter $z$. See Chapter 11 of \cite{Elvang:2013cua} for a review of 3d BCFW and its application in ABJM theory. The expression \reef{Is} is probably not the ideal form of the 6-point residue, since its expected properties are not manifest. For example, the result for the $n=6$ ABJM superamplitude should have $i \rightarrow i+2$ cyclic symmetry of the integral, as discussed in section \ref{s:orthog}; dressing \reef{Is} with the Jacobian from \reef{J234} should make it invariant under $i \rightarrow i+2$, but this is not obvious. Another point of concern about \reef{Is} are the apparent higher-order poles from the denominator-factors $\four{5612}^2$ and $\four{2356}^3$. The former is not too worrisome: using the identity \reef{4Dschouten} gives $\four{5612} = - \dtwo{51} \dtwo{62}= - \<56\> \tilde{E}_6 \<61\>^2 \tilde{E}_1 \<12\>$. Numerator factors can then cancel the individual angle brackets so that there are not double poles. The triple-pole at $\four{2356} =0$ would appear to be a worse problem because $\four{2356} = \<23\> \<56\> y_{63}^2 \propto (p_6+p_1+p_2)^2$, which is not just a simple product of angle-brackets. However, it is not hard to show that the numerator factors conspire to cancel the extra powers in $\four{2356}$ so that at most we have a simple pole at $\four{2356} =0$. A detailed argument is given in Appendix \ref{s:pole2356}. We leave further analysis of the momentum twistor form of the $n=6$ ABJM amplitude for future work. \subsubsection{Singularities of the residues} As a warm up, let us briefly review how poles of the NMHV residues could be understood as boundaries of cells in the $\mathcal{N}=4$ SYM Grassmannian. Choose a contour $\gamma_{abcde}$ such that the integral \reef{topC} picks up the residue where $n{-}5$ $c_i$'s with $i\ne a,b,c,d,e$ vanish. The remaining five $c_i$'s are then fixed by the four bosonic delta functions $\delta^4(c\cdot Z)$ and the $GL(1)$ scaling, as discussed in Section \ref{s:NMHVeval}; the result is the 5-bracket $[abcde]$. Its five poles can be described in the Grassmannian integral by forcing an extra $c_i$ to be zero (as discussed in Section \ref{s:bdr}): for example, if $c_a=0$, the delta-function $\delta^4(c\cdot Z)$ says that $Z_b$, $Z_c$, $Z_d$, and $Z_e$ are linearly dependent, and hence $\four{bcde} = 0$. This is precisely one of the five poles of the residue $[abcde]$. Now consider the ABJM momentum twistor space Grassmannian \reef{L61} for $n=6$. Since the integral is completely fixed by the bosonic delta functions and $GL(1)$, there is no contour to choose and the result is simply two terms that are conjugate to each other, one for each of the two branches determined by the orthogonal constraint. Let us gauge-fix the $GL(1)$ by setting $c_1=1$ and then analyze the constraints of the bosonic delta-functions when $c_2=0$. Via \reef{n6cGc} and \reef{n6orthocycl}, the orthogonality condition with $c_2=0$ can be written \begin{equation} c_5 \langle\!\langle 51 \rangle\!\rangle = c_4 \langle\!\langle 42 \rangle\!\rangle c_2 = 0\,. \label{orthoNow} \end{equation} So we must have $c_5=0$, or $\langle\!\langle 51 \rangle\!\rangle=0$. Examining each in turn: \begin{itemize} \item $c_2=c_5=0$: using $\Omega_{AB}$ to dot $Z_3$ and $Z_4$ into the constraint $c\cdot Z = 0$, we find \begin{equation} \langle\!\langle 13 \rangle\!\rangle + c_6 \langle\!\langle 63 \rangle\!\rangle = 0 ~~~~\text{and}~~~~ \langle\!\langle 14 \rangle\!\rangle + c_6 \langle\!\langle 64 \rangle\!\rangle = 0\,. \end{equation} For this to hold true with non-vanishing $c_6$ requires $\four{6134} = 0$. The relations in $c\cdot Z = 0$ can then be solved to find \begin{equation} c_3 = \frac{\langle\!\langle 51 \rangle\!\rangle}{\langle\!\langle 35 \rangle\!\rangle}\,, ~~~~~ c_4 = \frac{\four{6123}}{\four{2346}}\,, ~~~~~ c_6 = \frac{\langle\!\langle 13 \rangle\!\rangle}{\langle\!\langle 36 \rangle\!\rangle}\,. \end{equation} Thus $c_2=c_5=0$ leaves the four other $c_i$ non-zero but imposes the constraint $\four{6134} = 0$ on the external data. This is precisely one of the poles in \reef{Is}. \item $c_2=0$ and $\langle\!\langle 51 \rangle\!\rangle=0$. Dotting $Z_5$ into $c\cdot Z = 0$ now gives $c_3 \langle\!\langle 35 \rangle\!\rangle=0$. If $c_3 = 0$, then $c\cdot Z = 0$ gives $c_4 \langle\!\langle 41 \rangle\!\rangle=0$ and $c_4 \langle\!\langle 46 \rangle\!\rangle=0$. If also $c_4 = 0$ then we get a lower-dimensional subspace ($c_2=c_3=c_4=0$). If $c_4 \ne 0$, we must have $\langle\!\langle 41 \rangle\!\rangle = \langle\!\langle 46 \rangle\!\rangle = 0$ which in addition to $\langle\!\langle 51 \rangle\!\rangle=0$ renders multiple 4-brackets to be zero. If instead $c_3\ne 0$, we must have $\langle\!\langle 35 \rangle\!\rangle=0$. Consistency of $c\cdot Z = 0$ then requires $\four{6134}=0$ and this combined with $\langle\!\langle 51 \rangle\!\rangle=\langle\!\langle 35 \rangle\!\rangle=0$ puts several constraints on the external data. The remaining conditions in $c\cdot Z = 0$ do not completely fix the rest of the $c_i$'s. \end{itemize} We conclude from the above that the three `bounderies' $c_i = c_{i+3} = 0$ in the $n=6$ orthogonal Grassmannian integral \reef{L61} correspond to poles of the form $\four{i{+}1,i{+}2,i{+}4,i{+}5} = 0$. These are exactly the three different 3-particle poles $P^2_{i{+}2,i{+}3,i{+}4}$ of the amplitudes: for example $P_{123}^2 = y_{14}^2 \propto \four{6134}$. One the other hand, boundaries of the form $c_{i} = c_{i+1} =0$ impose constraints among two-brackets (e.g.~ $\langle\!\langle 51 \rangle\!\rangle=\langle\!\langle 41 \rangle\!\rangle = \langle\!\langle 46 \rangle\!\rangle =0$) which are akin to soft limits. Thus, as in the 4d case, we find that the cell boundaries of the Grassmannian correspond to poles in the residues. The fact that locality is partially hidden in the orthogonal constraint was already observed in the derivation of the twistor string formula for ABJM theory~\cite{Huang:2012vt},\footnote{A twistor string whose vertex operators give the corresponding formula was later presented in~\cite{Engelund:2014sqa}.} where locality is achieved as the momentum space Grassmannian is localized onto a Veronese map~\cite{ArkaniHamed:2009dg}, $B_{\alpha, i}(a_i,b_i)=a_i^{\tilde{k}-\alpha} b_i^{\alpha-1}$. This reduces a G$(k,n)$ down to a G$(2,n)$ Grassmannian, parameterized by $(a_i, b_i)$. At 6-point, this localization was achieved by the orthogonal constraint. For higher-points, only part of the orthogonal constraint is relevant to the localization to the Veronese map. Thus we anticipate that for higher-points, the effective metric will continue to play an important role for the realization of locality. \section{Outlook} \label{s:out} In this paper, we presented a detailed review of the relationship between Grassmannian integral formulas of different external data and provided a new derivation to establish their equivalence. Using the new approach, we derived the momentum twistor version of the Grassmannian integral for ABJM theory. Contrary to the momentum space representation, which is an orthogonal Grassmannian with constant metric, the momentum twistor space representation corresponds to an orthogonal Grassmannian whose metric depends on the external data. There are a number of interesting questions that can be tackled at this point. Recently the planar amplitudes of $\mathcal{N}=4$ SYM have been identified as a single geometric object, the amplituhedron~\cite{Arkani-Hamed:2013jha,Arkani-Hamed:2013kca}. It is defined in the Grassmannian G($k$,$k$+4) via \begin{equation}\label{Yspace} Y^I_{\alpha }=C_{+,\alpha i}Z^I_{+i}\,, \end{equation} where $C_{+,\alpha i}$ are cells in the positive Grassmannian\footnote{The positive Grassmannian, or non-negative Grassmannian to be precise, refers to the property that the $k \times n$ matrices are real-valued with all minors are greater or equal to zero.} G($k,n$) and $Z^{I}_{+,i}$ are $(k+4)$-component vectors, $i=1,\ldots,n$, built linearly from the momentum supertwistors (see~\cite{Arkani-Hamed:2013jha} for details). The array of $n$ vectors $Z^{I}_{+,i}$ are viewed as elements in the positive Grassmannian G($k{+}4,n$). The amplituhedron is then the ``volume-form" in this space, and it has logarithmic singularities at the boundaries of $Y$. It would be very appealing to derive this definition from the momentum twistor space Grassmannian integral. As a first step, one should be able to prove that the BCFW terms in momentum twistor space are associated with dimension $4k$ positive cells in G($k$,$n$). In principle, this is accomplished by the momentum twistor space on-shell diagrams introduced by He and Bai~\cite{HeBai}, where the individual cells are associated with diagrams that are again iteratively built from the fundamental 3-point vertices. On the other hand, from our analysis one might expect the existence of a straightforward map from cells in the momentum space Grassmannian to cells in the momentum twistor space Grassmannian. However as the minors of the two Grassmannians are related by a multiplicative string of spinor brackets, a priori it is not clear that positivity in one Grassmannian can be related to that of the other, even for the top-cell. Thus we see that positivity of the BCFW terms in momentum twistor space is non-trivial result, and should warrant further investigation. Given that we have derived the momentum twistor space Grassmannian for ABJM theory, one can ask if there exists a geometric entity like the amplituhedron for ABJM? Supporting evidence for its existence includes the realization that BCFW recursion relations for the theory exists both at tree- and at loop-level, and when represented in terms of on-shell diagrams, stratifies the positive orthogonal Grassmannian. Unlike $\mathcal{N}=4$ SYM, where positivity ensures locality, we have already seen that the orthogonal condition plays an important role as well. What is unclear is whether or not the condition is to be viewed as a condition on the cells, or on the space $Y$ where the amplituhedron lives, perhaps both. Another interesting question is if there exist on-shell diagrams for cells in ABJM momentum twistor space Grassmannian such as those found for $\mathcal{N}=4$ SYM~\cite{HeBai}. One of the remarkable results in the on-shell diagram approach for cells of the momentum space Grassmannian, is that the gluing and merging of diagrams preserves orthogonality~\cite{ArkaniHamed:2012nw,HW}. In momentum twistor space, the orthogonality is now defined with a momentum twistor dependent metric. It would be very interesting if an iterative way of constructing cells exists such that the orthogonality property of the smaller cells ensures that of the higher-dimenions ones. In this paper, we also studied residue theorems for the NMHV level in momentum twistor space and showed that an abstract homological point of view offered a clear geometric description. For amplitudes beyond NMHV, the combinatorics and geometry become much more difficult, and it is no longer the case that every momentum twistor cell of the appropriate dimension ($d=4k$) has support for generic external data. Composite singularities become the norm, even in momentum twistor space, so the residue calculation outlined in Appendix \ref{app:residues} cannot be applied directly. Moreover, whereas the entries in the $k=1$ matrix \eqref{topC} can be interpreted as homogeneous coordinates on $\mathbb{CP}^{n-1}$, higher $k$ Grassmannians do not have such simple geometric structure. Even the locations of poles become more complicated for $k>1$ due to the non-linear dependence of the minors on the matrix entries. Instead of cutting out hyperplanes as in Section \ref{s:resrels}, the minors vanish on generally complicated surfaces. Understanding the geometric and homological structure of such spaces in a general sense remains a subject of active research in the mathematics community. Despite the inherent mathematical challenges, the BCFW bridge decompositions of \cite{ArkaniHamed:2012nw} suggest a possible route forward. The technique provides a robust way to generate coordinates on any cell of G$(k,n)$ with the useful property that all singularities of the integration measure are manifestly of the form $d\alpha/\alpha$ (similar to the $dc/c$ structure of \eqref{NMHVlink}). This eliminates the issue of composite residues at the cost of requiring multiple charts to cover all singularities.\footnote{Until recently it was not known how to compare the orientations of those charts, but this has since been resolved \cite{OlsonInprep}.} The existence of such a convenient representation of the Grassmannian integral offers compelling motivation to pursue higher $k$ generalizations, and it will almost certainly lead to further insights regarding residues, residue theorems, superamplitudes, and locality constraints for all $k$ and $n$. \section*{Acknowledgements} We would like to thank Nima Arkani-Hamed, Jake Bourjaily, Freddy Cachazo, and Jaroslav Trnka for useful and insightful discussions. Y.-t.H.~would also like to think Dongmin Gang, Eunkyung Koh, Sangmin Lee, and Arthur E.~Lipstein for early collaborations on related subjects. H.E.~is supported in part by NSF CAREER Grant PHY-0953232. C.K.~is supported in part by the US Department of Energy under grant DE-FG02-95ER40899 and she would also like to thank ASU for hospitality during the final stages of this work. T.L.~is supported by NSF grant DMS-1160726. T.M.O.~is supported by NSF Graduate Research Fellowship under Grant \#F031543. S.B.R is supported in part by US Department of Energy under grant DE-SC0011719.
\section{Data description and results} We consider here a database on mixed marriages and newborns to mixed couples, annually recorded by ISTAT (Italian Institute of Statistics) for all italian municipalities in the time period from 2001 to 2011. In that time span documented foreign born population has grown in Italy from just over 1 million to almost 4 millions people, corresponding to an increase from $2\%$ to over $6\%$ of the total resident population. These figures do not include illegal immigrants whose numbers are difficult to determine. Although there are immigrants from almost all nations, the communities with higher presence are from Romania ($969$,$000$ in $2011$, i.e $21\%$ of the total population of immigrants), followed by Albania ($483$,$000$ in $2011$, i.e $11\%$), Morocco ($452$,$000$ in $2011$, i.e $10\%$), and China ($210$,$000$ in $2011$, i.e $5\%$). The geographical distribution of immigrants is largely not homogeneous: $86.5\%$ lives in the northern and central part of the country (the most economically developed areas), while $13.5\%$ lives in the south. The dataset contains over $1$,$100$,$000$ data, yearly describing - for each of the $8$,$100$ municipalities - the total population, the number of immigrants, the number of marriages and newborns originating from different types of couples (either mixed or not). For each municipality we have considered the immigrant density \begin{equation} \centering\gamma=\frac{N_{imm}}{N_{imm}+N_{nat}} \; , \end{equation} where $N_{imm}$ is the number of immigrants and $N_{nat}$ is the number of the natives. A useful parameter, measuring the number of possible cross-links among natives and immigrants is $\Gamma=\gamma(1-\gamma)$. We focussed, in particular, on two quantifiers: the fraction of marriages with spouses of mixed origin (native and immigrant) $M_m$ and the fraction of newborns with parents of mixed origin $B_m$. The entire marriages dataset contains $89$,$093$ records but only $82$,$208$ points were considered in our study. This is due to the fact that in tiny villages, occasionally, no marriages occurred. Those events account for about $8\%$ of the whole dataset. Moreover, in about $56\%$ of all records no mixed marriages have occurred. Analogously the inputs from municipalities where no newborns occurred account for about $4\%$ of the whole dataset, whereas for about $43\%$ of all records no mixed newborns have occurred. We separated the dataset into two parts: one for the small cities (below $10$,$000$ inhabitants) and the other for large ones (above $10$,$000$ inhabitants). Figure \ref{FigRaw} shows the resulting collection of the raw data of each part for the mixed marriages (upper panels) and for the newborns of mixed couples (lower panels) in the planes $(\gamma,M_m)$ and $(\gamma,B_m)$, respectively. We observe that records belonging to villages and small cities, whose population is under $10$,$000$ inhabitants, count up to $84\%$.\ Figure \ref{FigDens} displays the data densities and shows statistical robustness (up to one percentile) for all $\gamma$ up to $16\%$. The partition of both marriages and newborns datasets, which has led us to the main results of this work, has been investigated under various perspectives. We remark that this partition was proposed since an analysis performed over a unified datasets (small and large cities together) happened to be unsatisfactory. It displayed in fact a high dependence on the binning parameter settings revealing the typical presence of {\it data mixture} of interacting and non-interacting type. This situation led us to partition each dataset, separating small municipalities from large ones. The implemented threshold (i.e. $10$,$000$ inhabitants) was attempted according to previous work \cite{bcsv}, where the considered municipalities consisted only of cities over $10$,$000$ inhabitants. Ex-post an optimisation test on the threshold has been performed showing that the proposed one was a stable choice, i.e. the coefficients of determination ($R^2$) of the data fitting in the two regimes were maximal. Moreover, for the chosen partition we verify both robustness and homogeneity among the two datasets (large and small cities) in order to exclude pathologies by imbalance. Our tests show that the immigrant proportions over the total population in large and small cities are comparable for each year with a maximum difference between the two of about one percent. We can therefore infer that immigrants seem to be distributed in the same manner, independently from the size of the municipality they belong to. Eventually, the ratio between the total population living in small municipalities and the one living large municipalities is nearly constant ($\sim 47\%$) over the years. We tested, moreover, the behavior of the temporal series (year by year) of the proportions of immigrants, of natives and of the total population living in the small cities. This study shows that the national fraction of immigrants and natives in small municipalities are about the same ($30\%$). Furthermore, a similar value (slightly more than $30\%$) is found for the total population that lives in cities with less than $10$,$000$ inhabitants. Since our study is devoted to investigate the average behavior at the country scale for the two quantifiers $M_m$ and $B_m$ in large cities and small ones, we estimate the average percentage of mixed marriages or mixed newborns, for a given immigrant density. To this purpose we performed a {\it mediant} average and binning with constant information (see \cite{bcsv} for details). Figure \ref{FigRadLine} displays the output averages versus $\Gamma$. The emerging behaviors are well fitted (see the $R^2$ coefficient) by two different laws: square root for the quantifiers on small municipalities and linear on large ones. The statistical physics approach developed and broadly described in the technical paper \cite{bcsv} suggests an interpretation for these results. We know in fact that the linear behaviour emerges, in strong ties conditions, from collective effects when the network is sparse and unpercolated. The links are rare and the connected groups are made of only few units. On the other side, the square root law emerges when the social framework of strong personal ties, is built on a fully connected network where the cases of isolated individuals or small groups inside the communities are rare or completely absent. In order to better test the predictability power of our theory we have performed our analysis not only for whole time span of the observation data, but also on increasing time sub-spans. Figure \ref{FigR2} shows that the distinction among linear and square root growth is already evident from elaboration coming from only $2001$ data, and stay stable for increasing intervals $2001-2002$ etc, up to the whole set of data ($2001-2011$). Our results are in good agreement with the classical sociological theories of alienation and anomie about the social behavior on large cities, where social connections are seldom and ineffective, in comparison to villages where they are strong (see \cite{durkheim}). Our results moreover have the potential to impact policy makers. In fact, the identification of the growth law of each quantifier in $\gamma$ allows to predict the level of the quantifiers for growing values of $\gamma$.
\section{Introduction} Effective equations for quantum systems developed in~\cite{EffAc} and~\cite{EffEq2} allow quick access to the behavior of a quantum system in the semiclassical regime without performing a formal Hilbert space construction. The usefulness of this approach has been showcased for non-harmonic quantum systems~\cite{Boj_Brahm_Nels} and cosmological models~\cite{EffCosm}. However, this method was originally developed for the situations where the fundamental kinematical variables of the system to be analyzed form the so-called canonical algebra (sometimes called Weyl-algebra). In order to make it applicable to a more general class of cosmological models or, more ambitiously, to gauge theories, it needs to be extended to the situations, where the defining variables form a more general Lie algebra. The major new feature of such observable algebras is the presence of ``redundancy'' conditions expressed by Casimir operators. In~\cite{Bojowald_sl2c_eff} these redundancy conditions are imposed following the methods of ``effective quantum constraints'' developed in~\cite{EffCons} and~\cite{EffConsRel}. While, by itself such a procedure is sensible, it remained to be demonstrated that it can be consistently implemented alongside the semiclassical truncation of the system. Establishing the consistency of combining semiclassical truncation and removing redundancy is the main result of this report. We work in the setting of a general finite-dimensional Lie algebra and perform explicit computations for the case of a single non-trivial center-generating element. The discussion is organized into three sections. Section~\ref{sec:QLieAlg} describes our approach to states on a Lie algebra and sets up both the semiclassical truncation and the reduction by the central element, summarizing the detailed results from subsequent sections. Section~\ref{sec:PBOrder} establishes the consistency of the reduction in the degrees of freedom and the quantum Poisson structure that they inherit from the Lie algebra during the semiclassical truncation. Section~\ref{sec:Constraints} uses explicit counting to determine the conditions under which the correct number of redundancy conditions remain after the semiclassical truncation. \section{Semiclassical phase-space of a Lie algebra} \label{sec:QLieAlg} We will focus our attention on the case of an associative unital algebra $\mathcal{A}_{kin}$, which is the universal enveloping algebra of some Lie algebra $\mathfrak{a}$, such that the center of $\mathcal{A}_{kin}$ is generated by a single independent non--trivial element $\hat{P}$, for example, the Casimir element in the case of semisimple Lie algebras. The case where the center of $\mathcal{A}_{kin}$\ has several independent generators can be treated in a similar way, by considering one generator at a time. Since our technique has been developed to complement the usual representation-based methods of quantum mechanics, the space of all linear states on $\mathcal{A}_{kin}$\ is of particular interest. Although $\mathcal{A}_{kin}$\ is infinite-dimensional as a linear space, to a given semiclassical order its space of states can be captured by a finite set of (non-linear) functions with the original Lie algebra structure giving rise to a (typically degenerate) Poisson bracket. The degree of freedom redundancy associated with $\hat{P}$\ can be accounted for alongside this truncation in the form of polynomial conditions on the non-linear functions mentioned above. In the rest of this section we first give some general quantum-mechanical motivation for considering states on the quotient space $\mathcal{A}_{kin}/ \mathcal{A}_{kin} \hat{C}$, where $\hat{C}=\hat{P}-r\hat{\mathbf{1}}$; we refer to passing to this quotient as ``reduction''. We then briefly overview the quantum phase-space generated by linear states on $\mathcal{A}_{kin}$. In~\ref{sec:Semiclassical} we describe the semiclassical truncation on this quantum phase-space. We conclude this section by considering the conditions under which the semiclassical truncation and reduction in the degrees of freedom give consistent results. The details of our argument for consistency are presented in the sections that follow. \subsection{Quantum system as an algebra} Quantum description of a finite--dimensional mechanical system can be thought of as a particular irreducible representation of a chosen (finite--dimensional) Lie algebra of phase-space functions that describe freedoms of the system classically. For example, the classical degrees of freedom of a particle moving in one dimension can be captured by its position $x$ and momentum $p$, subject to the canonical Poisson relation $\{x, p\}=1$. Within the usual ``Schr\"odinger'' quantum mechanics of a particle in one dimension the degrees of freedom are captured by the canonical algebra defined by the commutation relation $[\hat{x}, \hat{p}] = i\hbar \hat{\mathbf{1}}$, where $\hat{x}$\ and $\hat{p}$\ are now differential operators, generating an infinite-dimensional associative algebra. The slightly more general situation of a Lie algebra with a central element in its corresponding universal enveloping algebra arises when quantizing a mechanical system with a phase-space $\Gamma$\ that possesses an over-complete set of coordinates $\{x_i\}$, $i=1, 2, \ldots M$, such that: \begin{enumerate}[(i)] \item $\{x_i\}$\ resolve the points of $\Gamma$; \item they form a Lie algebra under the Poisson bracket: ${\displaystyle \{x_i, x_j\} = \alpha_{ij}^{\ \ k} x_k}$, where $\alpha_{ij}^{\ \ k}$\ are structure constants (typically real) and summation over repeated indices is implied; \item the over-completeness is expressed by a single polynomial relationship $P(x_1, \ldots, x_M) = const.$, with $\{P, x_i\} = 0$ for all coordinate functions $x_i$. \end{enumerate} A simple example of such a system is the sphere $\Gamma \cong S^2$, with the area form providing the symplectic structure. Viewed as embedded in the Euclidean space $\mathbb{R}^3$\ it has an over-complete set of Cartesian coordinates $x$, $y$, $z$, that obey the $su(2)$\ Lie algebra (of rotations about the three coordinate axes) with respect to the corresponding Poisson bracket. The redundancy is expressed by the Casimir element of $su(2)$, we have $x^2+y^2+z^2 = const.$\ Another example is provided by the bouncing cosmological model introduced in~\cite{Bojowald_sl2c_1}~and~\cite{Bojowald_sl2c_2}, which is based on the Lie algebra $sl(2, \mathbb{C})$. More generally, our hope is to extend the effective method to the quantum treatment of gauge fields, whose degrees of freedom at each spatial point are captured by a finite--dimensional Lie algebra. By analogy with the standard notation for operators in quantum mechanics, we denote the generators of the universal enveloping algebra by $\hat{x}_i$, using ``\ $\widehat{\ }$ '' to distinguish elements of $\mathcal{A}_{kin}$\ emphasizing that multiplication is no longer abelian as it was in classical mechanics. The universal enveloping algebra is the collection of all (typically complex) polynomials in $\hat{x}_i$\ modulo the canonical commutation relations (CCRs)\footnote{The factor of $\hbar$\ appearing in the CCR can be treated as a formal parameter keeping track of the number of commutators taken. It does not change the Lie algebra structure and can be easily absorbed by using $\hat{\chi}_i = \hat{x}_i /\hbar$\ as generators.} \begin{equation} \label{eq:CCR} [\hat{x}_i, \hat{x}_j] := \hat{x}_i\hat{x}_j - \hat{x}_j\hat{x}_i = i\hbar \alpha_{ij}^{\ \ k} \hat{x}_k \ . \end{equation} This is one way of enforcing the so-called ``Dirac's condition'' for quantization, that establishes the correspondence between the quantum commutator and the classical Poisson bracket. Here $\hbar$\ introduces the quantum scale---we will later explicitly treat it as a small quantity for the purpose of semiclassical truncation. The universal enveloping algebra is Lie-homomorphic to any Hilbert space representation of such a system and therefore contains its representation-independent features. We will refer to $\mathcal{A}_{kin}$\ as the \emph{kinematical} algebra of the system because it describes the system before the redundacy conditions and its dynamical properties (e.g. selecting a prefered Hamiltonian element) are imposed. Here we will not worry about the dynamics, concentrating instead on the semiclassical method of imposing the redundancy conditions. Since any physical properties of interest must preserve the true phase-space defined by $P(x_1, \ldots, x_M) = const.$, we expect that the true quantum degrees of freedom are given by the quotient $\mathcal{A}_{kin}/\mathcal{A}_{kin} \hat{C}$, where $\hat{C} = \hat{P} - r\hat{\mathbf{1}}$, $\hat{P}$\ is the quantization of $P$\ and the generator of the center of $\mathcal{A}_{kin}$\ and $r$\ is some real-valued constant. In a semisimple Lie algebra, $\hat{C}$\ would fix the value of the Casimir charge. In reference to this class of systems we will refer to it as the \emph{Casimir constraint}\footnote{In the case of a compact phase-space defined by the condition $P(x_1, \ldots, x_M) = const.$, $r$\ in general cannot take arbitrary real values, but must be set to one of the allowed quantum numbers for consistency. However, in the semiclassical regime the phase space volume, and hence also $r$, must be sufficiently large compared to the gaps between their possible discrete values to allow existence of ``sharply peaked'' states. Thus in what follows the precise value $r$\ will not be important.}. For example, in the above example of a spherical phase-space, the Casimir constraint fixes the size of the phase space $\hat{C} = \hat{x}^2 + \hat{y}^2 + \hat{z}^2 - R^2\hat{\mathbf{1}}$. Since $\hat{C}$\ commutes with every element of $\mathcal{A}_{kin}$, the quotient is in fact a well-defined associative unital algebra. In the following subsections we briefly describe the setup for creating a semiclassically truncated version of this quotient algebra. \subsection{The quantum phase-space} \label{sec:QPhaseSpace} For our purposes, states on $\mathcal{A}_{kin}$\ are linear maps from $\mathcal{A}_{kin}$\ to complex numbers. The space of all such maps that, in addition, assign 1 to the identity element $\hat{\mathbf{1}}$, will be referred to as the quantum phase-space $\Gamma_Q$\ of the system. \footnote{In algebraic constructions of representations one typically imposes the positivity condition on states, however in the case of $\Gamma_Q$\ we make no such restriction to begin with. In particular applications of the effective technique positivity conditions are often imposed once constraints on $\mathcal{A}_{kin}$ are accounted for. In the language of constrained quantum mechanics, the states on the full algebra $\mathcal{A}_{kin}$\ can be thought of as kinematical.} By the natural duality, each element of $\mathcal{A}_{kin}$\ is also a linear function on $\Gamma_Q$. We will denote the function induced by $\hat{a} \in \mathcal{A}_{kin}$\ as $\langle \hat{a} \rangle$\ by direct analogy with ordinary quantum mechanics, where the expectation value of an operator is a specific example of such a function. Indeed, we will explicitly refer to functions of the type $\langle \hat{a} \rangle$\ as \emph{expectation values}. Evidently, if $\varphi$\ is an element of $\Gamma_Q$, then there is a complex number $\varphi(\hat{a})$\ that defines the value taken by the expectation value function $\langle \hat{a} \rangle$\ at the quantum phase-space point $\varphi$. It is straightforward to establish that this identification between the elements of $\mathcal{A}_{kin}$\ and expectation value functions on $\Gamma_Q$\ is linear, e.g. $\langle \hat{a} + \hat{b} \rangle = \langle \hat{a} \rangle + \langle \hat{b} \rangle$. Additionally, the expectation value is ``normalized\rq{}\rq{}, that is $\langle \hat{\boldsymbol{1}}\rangle = 1 \ $. Importantly, a subset of expectation value functions resolves the points of $\Gamma_Q$. This follows as $\mathcal{A}_{kin}$\ has a (countable) linear basis, for example polynomials of the form $\hat{x}_1^{i_1}\hat{x}_2^{i_2} \ldots \hat{x}_M^{i_M}$, where $M$\ is the dimension of the Lie algebra. Therefore, an element of $\Gamma_Q$\ is completely defined by the values it assigns to such a basis, meaning that the expectation value functions induced by any such basis form a complete set of coordinates on $\Gamma_Q$. Since $\Gamma_Q$\ is the (vector space) dual of an algebra, it possesses additional structure. Most importantly, the Lie-algebraic structure inherited by $\mathcal{A}_{kin}$\ from the original Lie algebra $\mathfrak{a}$ is passed along to $\Gamma_Q$\ in the form of the quantum Poisson bracket $\{.,.\}_Q$. For a pair of expectation value functions it is defined directly by \begin{equation}\label{eq:QPoisson} \left\{ \langle \hat{a}\rangle ,\langle \hat{b} \rangle \right\}_Q := \frac{1}{i\hbar} \left\langle [\hat{a}, \hat{b} ] \right\rangle = \frac{1}{i\hbar} \left\langle \hat{a} \hat{b} - \hat{b} \hat{a} \right\rangle \ . \end{equation} Jacobi identity for the above bracket follows quickly from properties of the commutator in an associative algebra. The bracket extends to non-linear functions on $\Gamma_Q$\ by linearity and Leibniz rule. The quantum Poisson bracket allows one to formulate dynamical relations of quantum mechanics in the form closely analogous to the classical Hamiltonian mechanics, which is the original motivation for developing this approach. In particular, the unitary time-evolution of a quantum wave-function, leads to the standard result of quantum mechanics for the corresponding time-evolution of the expectation value of any (time-independent) operator \[ \frac{d}{dt} \langle \hat{a} \rangle = \frac{1}{i\hbar} \left\langle [\hat{a}, \hat{H} ] \right\rangle \ , \] where $\hat{H}$\ is the Hamiltonian of the system. In terms of the quantum Poisson bracket, this is simply \begin{equation} \label{eq:Qevolution} \frac{d}{dt} \langle \hat{a} \rangle = \left\{ \langle \hat{a}\rangle ,\langle \hat{H} \rangle \right\}_Q \end{equation} By linearity and Leibniz rule, this result extends to any function $f$ on $\Gamma_Q$, so that $df/dt = \{f, H_Q \}_Q$, where $H_Q = \langle \hat{H} \rangle$\ is refered to as the \emph{quantum Hamiltonian} in the literature. Since the degrees of freedom of $\mathcal{A}_{kin}$\ contain redundancy, so does the quantum phase-space $\Gamma_Q$. Algebraic redundancy is (formally) removed by passing to the quotient $\mathcal{A}_{kin}/\mathcal{A}_{kin} \hat{C}$. The true quantum phase-space should therefore be the space of linear states on the above quotient algebra. This can be implemented by passing to the subspace of $\Gamma_Q$\ that vanishes on $\mathcal{A}_{kin} \hat{C}$. In particular, ``physical'' states must satisfy \[ \langle \hat{a} \hat{C} \rangle = 0 \quad , \quad \forall \hat{a} \in \mathcal{A}_{kin} \ . \] This can be systematically implemented through a (countably) infinite set of conditions using any basis on $\mathcal{A}_{kin}$, for example \begin{equation}\label{eq:CPoly} \left\langle \hat{x}_1^{i_1}\hat{x}_2^{i_2} \ldots \hat{x}_M^{i_M} \hat{C} \right\rangle = 0 \quad , \quad \forall \ \ i_1, i_2, \ldots i_M \ . \end{equation} Consider $\Sigma$, the linear subspace of the phase space $\Gamma_Q$, where all of the above conditions are satisfied. We quickly infer that it is a Poisson submanifold of $\Gamma_Q$, since \[ \left\{ \langle \hat{a} \rangle, \langle \hat{b} \hat{C} \rangle \right\}_Q = \frac{1}{i \hbar}\left\langle \left[ \hat{a}, \hat{b} \right] \hat{C} \right\rangle =_{\Sigma} 0 \ , \] meaning that all Poisson--generated vector fields are tangent to $\Sigma$, as well as that the constraint functions themselves generate no flows on $\Sigma$\ at all. In other words, $\Sigma$\ naturally inherits the Poisson structure from $\Gamma_Q$. \subsection{Semiclassical hierarchy} \label{sec:Semiclassical} Although, strictly speaking, the quantum phase space $\Gamma_Q$\ was constructed as a linear (or, more accurately, \emph{affine}) space, we will want to think of it as a differential manifold. More precisely, we think of it as an extension of the classical phase space $\Gamma$. The latter can be identified with $M$-dimensional subspaces of $\Gamma_Q$, formed by allowing the expectation values of the Lie algebra generators to vary, while keeping all other ``coordinates'' constant. We will designate one such subspace as being ``maximally classical'': semiclassical approximation will be valid in some small neighborhood of this subspace. A completely specified (non-distributional) classical state assigns values in accordance with \[ \varphi(f(x_i)) = f(\varphi(x_i)) \ , \] for any function $f$. So that, e.g. the value of distance squared is the square of the value of distance and so on. We can express these ``maximal classicality'' conditions for $\Gamma_Q$\ most naturally using polynomials \begin{equation} \label{eq:max_class} \left\langle \widehat{pol} \right\rangle = pol(\langle \hat{x}_1 \rangle, \langle \hat{x}_2 \rangle, \ldots ) \ , \end{equation} where $pol$\ is some function polynomial in the classical coordinate functions and $\widehat{pol}\in\mathcal{A}_{kin}$\ is the corresponding element of the algebra of quantum operators. We immediately run into difficulties, as there is no unique natural identification between classical polynomial functions and elements of $\mathcal{A}_{kin}$, since different choices of ordering generally yield distinct elements. For example, $\hat{x}_1 \hat{x}_2 \hat{x}_1$\ and $(\hat{x}_1^2\hat{x}_2+\hat{x}_2\hat{x}_1^2)/2$\ are both sensible\footnote{In quantum mechanics, we also typically want physical observables to correspond to self-adjoint operators, which should then be constructed out of *-invariant combinations of the basic generators. Typically (though not always) the chosen classical Poisson algebra is real, so that the the *-structure of the enveloping algebra is defined by $\hat{x}_i^* = \hat{x}_i$, and symmetric products of the generators give *-invariant elements of $\mathcal{A}_{kin}$. Hence the choice of symmetric orderings in this example and thereafter.} choices for the quantum analogue of $x_1^2x_2$, while, in general they are not the same. A brief calculation using~(\ref{eq:CCR}) yields \[ \hat{x}_1 \hat{x}_2 \hat{x}_1 = \frac{1}{2}(\hat{x}_1^2\hat{x}_2+\hat{x}_2\hat{x}_1^2) - \frac{\hbar^2}{2} \alpha_{21}^{\ \ i} \alpha_{1i}^{\ \ j} \hat{x}_j \ . \] In general, the second term on the right-hand-side does not vanish, however the expectation values it yields are ``comparatively small'', since it is suppressed by the factor of $\hbar^2$. In general, for any two choices of ordering for $\widehat{pol}$, the difference between the corresponding expectation values $\left\langle \widehat{pol} \right\rangle$\ is suppressed by $\hbar$\ (for a pair of symmetric orderings, the suppression is at least $\hbar^2$\ as in the above example, since at least two re-orderings are required). For this reason, it does not matter which ordering we choose when enforcing ``maximal classicality''~(\ref{eq:max_class}), since all choices will define classical submanifolds of $\Gamma_Q$\ that lie ``close'' to each other and will therefore define the same semiclassical neighborhood. We follow the literature on canonical effective equations~\cite{EffAc} and~\cite{EffEq2}, and identify $\widehat{pol}$\ with the totally-symmetrized polynomials in $\hat{x}_i$, we will refer to this as \emph{Weyl ordering}. In the above example this corresponds to \[ \widehat{x_1^2x_2} := \frac{1}{3} \left( \hat{x}_1^2\hat{x}_2 + \hat{x}_1\hat{x}_2\hat{x}_1 + \hat{x}_2\hat{x}_1^2 \right) =: \left( \hat{x}_1^2 \hat{x}_2 \right)_{\rm Weyl} \ . \] The discrepancy between polynomials in expectation values $\langle {\hat{x}_i}\rangle$\ and the expectation values of symmetrized polynomials are well captured by the so-called \emph{generalized moments} of the basic generators \begin{equation} \label{eq:moments} \Delta(x_1^{i_1} x_2^{i_2} \ldots ) := \left\langle \left( \hat{x}_1 - \langle \hat{x}_1 \rangle \right)^{i_1} \left( \hat{x}_2 - \langle \hat{x}_2 \rangle \right)^{i_2} \ldots \right\rangle_{\rm Weyl} \ . \end{equation} Note that $\Delta (x_i) = \langle \hat{x}_i -\langle \hat{x}_i \rangle \rangle = 0$. These moments are essentially non-linear functions on $\Gamma_Q$. Upon careful inspection, it is clear that the expectation value of any polynomial element of $\mathcal{A}_{kin}$ can be expressed in terms of the generalized moments and the expectation values of generators $\langle \hat{x}_i \rangle$. Together they provide an alternative (to expectation value functions) coordinate basis on $\Gamma_Q$. In particular, by writing $\hat{x}_i = \left( \hat{x}_i - \langle \hat{x}_i \rangle \right) + \langle \hat{x}_i \rangle$\ and expanding the product, it is not difficult to convince oneself that \begin{eqnarray} \label{eq:monomial_exp} \left\langle \hat{x}_1^{i_1} \hat{x}_2^{i_2}\ldots \right\rangle_{\rm Weyl} = & & (\langle \hat{x}_1\rangle^{i_1} \langle \hat{x}_2\rangle^{i_2} \ldots ) \\ \nonumber &+& \left. \sum_{n_1=0}^{i_1} \sum_{n_2=0}^{i_2} \cdots \frac{1}{n_1! n_2! \ldots} \ \frac{\partial^{\sum n_j} (x_1^{i_1} x_2^{i_2} \ldots )}{\partial^{n_1}x_1 \partial^{n_2} x_2 \ldots } \right|_{x_i=\langle \hat{x}_i \rangle} \Delta(x_1^{n_1} x_2^{n_2} \ldots ) \ , \end{eqnarray} where we set $\Delta(x_1^{0} x_2^{0} \ldots ) :=0$. It follows, that condition~(\ref{eq:max_class}) with Weyl--symmetrized choice for $\widehat{pol}$\ is satisfied precisely when all moments are set to zero. Therefore, $\langle \hat{x}_i \rangle$\ and $\Delta(x_1^{i_1} x_2^{i_2} \ldots )$\ can be thought of as the classical and quantum coordinates on $\Gamma_Q$\ respectively; we will refer to this particular choice of coordinate functions on $\Gamma_Q$\ as \emph{quantum variables}. We will refer to the sum $\sum_{n=1}^M i_n$\ as the \emph{order} of the generalized moment $\Delta(x_1^{i_1} x_2^{i_2} \ldots )$. A \emph{semiclassical} state is one that is ``sharply peaked'' about some values of a classically-complete set of observables, meaning, for example, that the standard deviations of these observables are small in such a state. For our purposes, a state on $\mathcal{A}_{kin}$\ is semiclassical, or, as we will more commonly say, the corresponding point lies in the \emph{semiclassical region} of $\Gamma_Q$, if the value it assigns to all moments of order $N$\ is comparable to the value of $\hbar^{N/2}$. This comparison assumes that all classical coordinates have been rescaled so that they have the same units as $\sqrt{\hbar}$, and can be explicitly realized in ordinary quantum mechanics using Gaussian states. An entirely equivalent definition of the semiclassical region would be obtained if we had chosen a different ordering when defining $\widehat{pol}$\ and moments in~(\ref{eq:moments}), since, because of the CCRs~(\ref{eq:CCR}) differently ordered products are equivalent up to terms proportional to powers of $\hbar$. We can therefore treat the semiclassical region of $\Gamma_Q$\ as well-defined by the choice of the generators $\hat{x}_i$\ alone and independent from our decision to work with the totally symmetrized products. When treating a physical quantum system, in a state that lies deep within the semiclassical region, one can safely neglect high-order moments that are suppressed by many powers of $\hbar$. In this situation, one can work with a finite-dimensional \emph{truncated} system, where all moments beyond a certain order have been discarded. This is the original motivation for studying the semiclassical hierarchy and defining and studying truncated systems. \subsection{Consistent truncation}\label{sec:Consistent_Trunc} We define $\Gamma_Q^{(N)}$, the truncated order $N$\ semiclassical state-space of the Lie algebra $\mathfrak{a}$, as the space with coordinates $\langle \hat{x}_i \rangle$\ and $\Delta(x_1^{i_1} x_2^{i_2} \ldots )$\ up to order $N$. In essence, we will identify $\Gamma_Q^{(N)}$ with the submanifold of $\Gamma_Q$\ defined by setting all moments beyond order $N$\ to zero. However, the quantum Poisson structure~(\ref{eq:QPoisson}) and the constraint conditions~(\ref{eq:CPoly}) need to be projected to $\Gamma_Q^{(N)}$\ with some care. Let us first define a more general notion of a semiclassical order: \begin{eqnarray} &(i)& \ {\rm Order } \left( \langle \hat{x}_i \rangle \right) = 0 \ , \label{eq:orderI} \\ &(ii)& \ {\rm Order\, } \left( \Delta(x_1^{i_1} x_2^{i_2} \ldots ) \right) = \sum_{n=1}^M i_n \ , \label{eq:orderII}\\ &(iii)& \ {\rm Order\, } \left( \hbar^{n} \right) = 2n \ , \label{eq:orderIII} \end{eqnarray} which we extend to monomials in basic expectation values, generalized moments, and $\hbar$\ by \[ {\rm Order\, } \left(f g \right) = {\rm Order\, } \left( f \right) + {\rm Order\, } \left( g \right) . \] A polynomial in $\hbar$\ and quantum variables of course generally mixes terms of different orders and does not itself posses a well--defined semicalssical order in the above sense however each of its monomial terms does. The notion of the order can be extended to polynomials as the \emph{leading semiclassical order} by defining Order $(.)$\ for a polynomial in $\hbar$\ and quantum variables as the lowest order of its non-zero monomial terms. It follows that \begin{equation*} {\rm Order\, } \left(f + g \right) \geq \inf \left\{ {\rm Order\, } \left( f \right), \, {\rm Order\, } \left( g \right) \right\} \ . \end{equation*} We define $N$--th order truncation operation on a monomial in the natural way \begin{equation} \label{eq:Trunc1} {\rm Trunc}_{(N)} (f) = \left\{ \begin{array}{l} f \ , \ {\rm if \ } {\rm Order\, } (f) \leq N \ , \\ 0 \ , \ {\rm if \ } {\rm Order\, } (f) > N \ . \end{array} \right. \end{equation} We extend it to polynomials by demanding linearity \begin{equation} \label{eq:Trunc2} {\rm Trunc}_{(N)} (f + g) = {\rm Trunc}_{(N)} (f) +{\rm Trunc}_{(N)} (g) \ . \end{equation} The truncation then simply cuts off all the terms of semiclassical order larger than $N$\ in the polynomial sum. The quantum Poisson bracket on $\Gamma_Q$\ consistently descends to a Lie bracket between a pair of order--$N$ truncated polynomials in a natural way \begin{equation} \label{eq:TruncatedPB} \{ f, g \}_Q^{(N)} := {\rm Trunc}_{(N)} \left( \{ f, g\}_Q \right) \ . \end{equation} In section~\ref{sec:PBOrder} we prove that \begin{equation} \label{eq:TruncPBconsistency} \{ f, g \}_Q^{(N)} = \{{\rm Trunc}_{(N)}( f) , \, {\rm Trunc}_{(N)}(g) \}_Q^{(N)} \ . \end{equation} ensuring that the order $N$\ truncation is a Lie algebra homomorphism from polynomials in $\hbar$\ and quantum variables of arbitrary order, to those of order $N$\ and below. This result also has an important practical implication for using a truncated description of a dynamical quantum system. In particular, this implies that the truncated quantum Hamiltonian, ${\rm Trunc}_{(N)}( H_Q)$, can be used to compute the order $N$\ evolution of the truncated system, considerably simplifying the computations involved in equation~(\ref{eq:Qevolution}). As we explicitly prove in section~\ref{sec:PBOrder} the above consistency condition is a special consequence of a general relation true for a pair of polynomials \begin{equation} \label{eq:PBorder} {\rm Order\, } \left(\{f, g\}_Q \right) \geq {\rm Order\, } \left( f \right) + {\rm Order\, } \left( g \right) -2 \ . \end{equation} Unfortunately, the truncation operation is not compatible with the ordinary abelian multiplication of polynomials, since in general \[ {\rm Trunc}_{(N)}( fg) \neq {\rm Trunc}_{(N)}( f){\rm Trunc}_{(N)}( g) \ . \] For example, $ {\rm Trunc}_{(3)} \left( \hbar \left\langle \left( \hat{x}_1 - \langle \hat{x}_1 \rangle \right)^2 \right\rangle \right) = 0$, while, since both factors are of order 2 \[ {\rm Trunc}_{(3)}\left( \hbar \right) {\rm Trunc}_{(3)} \left( \left\langle \left( \hat{x}_1 - \langle \hat{x}_1 \rangle \right)^2 \right\rangle\right) = \hbar \left\langle \left( \hat{x}_1 - \langle \hat{x}_1 \rangle \right)^2 \right\rangle \ . \] It follows that the truncated bracket defined by~(\ref{eq:TruncatedPB}) does not satisfy the Leibniz rule and is therefore not a true Poisson bracket.\footnote{The situation could be remedied by defining a truncated abelian product between polynomials along the lines of~(\ref{eq:TruncatedPB}).} With the degrees of freedom and the Poisson structure consistently truncated to order $N$, we are left with truncating the countably infinite set of Casimir constraint conditions given e.g. by equation~(\ref{eq:CPoly}). To proceed we first express the constraints using the expectation values of basic operators and their generalized moments, which is in principle always possible---one can imagine first symmetrically reordering polynomial terms within $\left(\hat{x}_1^{i_1}\hat{x}_2^{i_2} \ldots \hat{x}_M^{i_M}\hat{C}\right)$\ and then using the expansion in equation~(\ref{eq:monomial_exp}). At this stage, it seems that applying the truncation operation of~(\ref{eq:Trunc1}) and~(\ref{eq:Trunc2}) to the constraint conditions directly will do the trick, however this leads to inconsistent results~\cite{EffCons} and generally to an incorrect reduction in the degrees of freedom as we will see in detail in section~\ref{sec:Constraints}. Instead, following~\cite{EffCons} and~\cite{EffConsRel} we formally assign the ``classical\rq{}\rq{} constraint function evaluated on the expectation values of the generators $C(\langle \hat{x}_1 \rangle, \langle \hat{x}_2 \rangle , \ldots , \langle \hat{x}_M \rangle)$\ semiclassical order of 2 (i.e. same as $\hbar$) even though the expression involves no generalized moments or explicit factors of $\hbar$. While ultimately this particular order assignment is required for consistency, it can be physically motivated by noting that in the classical limit $C$\ must vanish on states that satisfy the Casimir constraint. In other words, in addition to the conditions~(\ref{eq:orderI})--(\ref{eq:orderIII}), when truncating the system of constraints we impose \begin{equation} \label{eq:orderIV} (iv) \ \ \ {\rm Order } \left( C(\langle \hat{x}_1 \rangle, \langle \hat{x}_2 \rangle , \ldots , \langle \hat{x}_M \rangle) \right) = 2 \ . \end{equation} In what way is the resulting truncated system of constraints \emph{consistent}? As is shown explicitly in section~\ref{sec:Constraints}, following this method of truncation we generate a finite number of non-trivial constraint conditions at each order. Moreover, provided that the non-trivial truncated constraints are functionally independent, the number of free degrees of freedom left in $\Gamma_Q^{(N)}$\ is the same as that of a system with no constraints that has one fewer basic generator. That is, provided certain regularity conditions are satisfied, the truncated tower of constraint conditions on expectation values and moments removes exactly one generating degree of freedom at each order. This is then our order $N$\ truncated phase space of the Lie algebra $\mathfrak{a}$: \begin{quote} The subspace $\Sigma^{(N)}$\ of the truncated space $\Gamma_Q^{(N)}$\ defined by setting the truncated set of constraint conditions $\{ {\rm Trunc}_{(N)} \left( \langle \hat{a} \hat{C} \rangle \right) : \hat{a} \in \mathcal{A}_{kin} \}$\ to zero, and equipped with a Lie bracket between polynomials in quantum variables given by~(\ref{eq:TruncatedPB}). The latter can be used to study truncated dynamics generated by a Hamiltonian or the truncated action of a symmetry group. \end{quote} In this section we formally defined the semiclassical truncation of states on a Lie algebra with a single center--generating (Casimir) element, generalizing the semiclassical states on the canonical algebra studied in~\cite{EffAc} and~\cite{EffEq2}. This setup has already been employed to study an $sl(2, \mathds{C})$--based quantum cosmological model in~\cite{Casimir} and we anticipate further applications to the development of effective formalism for gauge field theories where degrees of freedom at each spatial point are typically captured by a Lie algebra. In the following section, we cast the discussion of semiclassical hierarchy in algebraic form and use it to prove property~(\ref{eq:PBorder}) and the consistency result~(\ref{eq:TruncPBconsistency}). In section~\ref{sec:Constraints} we use this construction to find the conditions under which the truncated system of constraints correctly reduces the number of degrees of freedom. \section{Extended algebra and the quantum Poisson bracket} \label{sec:PBOrder} It is possible to define semiclassical hierarchy directly at the algebraic level rather than on the space of states, however the universal enveloping algebra $\mathcal{A}_{kin}$\ has to be extended for this purpose in order to accommodate the moment--generatig elements. We first define the \emph{classical} polynomial algebra corresponding to the Lie algebra $\mathfrak{a}$ as the commutative algebra of complex polynomials in the basic generators: $\mathcal{A}_{class} := \mathds{C}[x_1, x_2, \ldots x_M]$. When Passing from the algebraic picture to states, we will identify these ``classical variables\rq{}\rq{} $x_i$\ with the expectation values of the generators $\langle \hat{x}_i\rangle$. The extended quantum algebra is then linearly generated by finite sums of elements of $\mathcal{A}_{kin}$\ with multiplicative coefficients allowed to take values in $\mathcal{A}_{class}$. In other words $\mathcal{A}_{ext} := \mathcal{A}_{kin} \otimes \mathcal{A}_{class}\ $, and it can be algebraically generated by complex polynomials in elements of the form $\hat{x}_i$, $x_i\hat{\mathbf{1}}$, and $\hbar \hat{\mathbf{1}}$. With products between $\hat{x}_i$\ and $\hat{x}_j$\ governed by the CCRs while those between $x_i$\ and $x_j$\ as well as those between $\hat{x}_i$\ and $x_{j}$\ abelian (i.e. commutative). To emphasize that the associative product of $\mathcal{A}_{ext}$\ is non--commutative, we will use ``hats\rq{}\rq{} to denote its generic elements, $\hat{a} \in \mathcal{A}_{ext}$, just as we did with $\mathcal{A}_{kin}$. Of particular interest to the study of semiclassical states are elements of the form $\hat{x}_i - x_i\hat{\mathbf{1}}$, which will serve as the algebraic analogue of the generalized moments. We define \[ \widehat{\Delta x}_i :=\hat{x}_i - x_i\hat{\mathbf{1}} \ . \] Evidently, since $\hat{x}_i = \widehat{\Delta x}_i + x_i\hat{\mathbf{1}}$, the elements $\widehat{\Delta x}_i$\ form an alternative set of generators for $\mathcal{A}_{ext}$\ over $\mathcal{A}_{class}$. The extended algebra does possess a natural semiclassical order. As before, let us first define it for the preferred set of generators: \begin{eqnarray} &(i)& \ {\rm Order } \left( x_i \hat{\mathbf{1}} \right) = 0 \ , \label{eq:AorderI} \\ &(ii)& \ {\rm Order\, } \left( \widehat{\Delta x}_i \right) =1 \ , \label{eq:AorderII}\\ &(iii)& \ {\rm Order\, } \left( \hbar \hat{\mathbf{1}} \right) = 2 \ . \label{eq:AorderIII} \end{eqnarray} Once again, we extend the definition to monomials via \begin{equation} {\rm Order\, } \left(\hat{a} \hat{b} \right) = {\rm Order\, } ( \hat{a} ) + {\rm Order\, } ( \hat{b} )\ . \label{eq:OrderP1} \end{equation} In order to extend the definition of order to polynomials, we need to consider the effect of the CCRs~(\ref{eq:CCR}). Distinct monomials are not all linearly independent from each other, for example, $\widehat{\Delta x}_i \widehat{\Delta x}_j$\ is assigned Order $=2$\ by above definition, while, using CCR we can express this product as \begin{eqnarray} \widehat{\Delta x}_i \widehat{\Delta x}_j &=& \widehat{\Delta x}_j \widehat{\Delta x}_i + \left[ \hat{x}_i - x_i\hat{\mathbf{1}}, \ \hat{x}_j - x_j\hat{\mathbf{1}} \right] \nonumber \\ \label{eq:Reordering} &=& \widehat{\Delta x}_j \widehat{\Delta x}_i + \left[ \hat{x}_i, \ \hat{x}_j \right] \\ &=& \widehat{\Delta x}_j \widehat{\Delta x}_i + i\hbar \alpha_{ij}^{\ \ k} \hat{x}_k = \widehat{\Delta x}_j \widehat{\Delta x}_i + i\hbar \alpha_{ij}^{\ \ k} x_k \hat{\mathbf{1}}+ i\hbar \alpha_{ij}^{\ \ k} \widehat{\Delta x}_k \ , \nonumber \end{eqnarray} where the monomial terms in the final expression have orders 2, 2, and 3 respectively. We can make the definition consistent by assigning complex polynomials in $ x_i \hat{\mathbf{1}}$, $\widehat{\Delta x}_i$, and $\hbar \hat{\mathbf{1}}$\ the lowest order of all of the orders of its non-zero monomial terms expanded in any given basis (i.e. a particular ordering is chosen for the factors $\widehat{\Delta x}_i$). The above Order (.) operation on $\mathcal{A}_{ext}$\ can then be interpreted as the \emph{leading order} of a polynomial. Generalizing~(\ref{eq:Reordering}), two differently ordered monomials are equivalent up to adding terms of the same semiclassical order or higher, it follows that every element of $\mathcal{A}_{ext}$\ has a unique semiclassical order in the above sense. Some useful properties follow \begin{eqnarray} {\rm Order\, } \left(\hat{a} \hat{b} \right) &=& {\rm Order\, } \left( \hat{b} \hat{a} \right) \ , \label{eq:OrderP2}\\ {\rm Order\, } \left(\hat{a} + \hat{b} \right) &\geq& \inf \left\{ {\rm Order\, } ( \hat{a} ), \, {\rm Order\, } ( \hat{b} ) \right\} \ . \label{eq:OrderP3} \end{eqnarray} From~(\ref{eq:OrderP1}), (\ref{eq:OrderP2}) and~(\ref{eq:OrderP3}) it quickly follows that \begin{equation} {\rm Order\, } \left( \left[\hat{a}, \hat{b} \right] \right) \geq {\rm Order\, } ( \hat{a} ) + {\rm Order\, } ( \hat{b} ) \ . \label{eq:OrderP4} \end{equation} By virtue of the factor of $\hbar$\ appearing in the CCRs the commutator is also an ``order raising'' operation. This can be seen by considering what happens to the generators of $\mathcal{A}_{ext}$\ when commutators are taken: $[\hat{x}_i, \hbar\hat{\mathbf{1}} ] =[\hat{x}_i, x_j \hat{\mathbf{1}} ] = 0$, while $\left[ \hat{x}_i, \widehat{\Delta x}_j \right] = i\hbar \alpha_{ij}^{\ \ k} x_k \hat{\mathbf{1}} + i\hbar \alpha_{ij}^{\ \ k} \widehat{\Delta x}_k$. In each case the order is raised by at least one relative to the second argument of the commutator. By the property of the commutator of products it quickly follows that \begin{equation} {\rm Order\, } \left( \left[\hat{x}_i, \hat{a} \right] \right) \geq {\rm Order\, } ( \hat{a} ) + 1 \ , \end{equation} and, more generally, that \begin{equation} {\rm Order\, } \left( \left[\hat{a}, \hat{b} \right] \right) \geq \sup \left\{ {\rm Order\, } ( \hat{a} ), \, {\rm Order\, } ( \hat{b} ) \right\} + 1 \ . \label{eq:OrderP5} \end{equation} Typically~(\ref{eq:OrderP4}) provides a stronger lower bound on the semiclassical order of a commutator $[\hat{a}, \hat{b}]$. However if the order of either $\hat{a}$\ or $\hat{b}$\ is zero, then condition~(\ref{eq:OrderP5}) is stronger. We now establish the connection between the extended algebra and functions on the quantum phase space by defining the map $\langle . \rangle: \mathcal{A}_{ext} \rightarrow \mathscr{F}\left( \Gamma_Q\right)$, such that for any $\hat{a} \in \mathcal{A}_{kin}$\ and $f \in \mathcal{A}_{class}$ \begin{eqnarray} \label{eq:ExpVal} \left\langle f(x_1, x_2, \ldots) \hat{a} \right\rangle := f(\langle \hat{x}_1 \rangle, \langle \hat{x}_2, \rangle \ldots ) \langle \hat{a} \rangle \ , \end{eqnarray} where the $\langle . \rangle$\ operation on the right is performed as in Section~\ref{sec:QPhaseSpace}. The map extends to the rest of $\mathcal{A}_{ext}$\ by requiring it to be complex--linear and has a non-trivial kernel with $\langle \hat{a} \rangle = 0$\ for any element of the form $\hat{a} = \hbar^n f \widehat{\Delta x}_i$\ and linear combinations of such elements, since $\left\langle \widehat{\Delta x}_i \right \rangle = \langle \hat{x}_i \rangle - \langle x_i \rangle = 0$. This map is compatible with the definition of the leading semiclassical order in both spaces. That is \begin{equation} \label{eq:OrderPres} {\rm Order }\ \left( \langle \hat{f} \rangle \right) \geq {\rm Order }\ \left( \hat{f} \right) \ . \end{equation} To see this consider a symmetrized monomial $\hat{a} = f \hbar^n \left( \widehat{\Delta x}_1^{n_1} \widehat{\Delta x}_2^{n_2} \ldots \right)_{\rm Weyl}$, where $f \in \mathcal{A}_{class}$. By definition~(\ref{eq:AorderI})--(\ref{eq:AorderIII}) Order\,$(\hat{a}) = 2n + \sum n_i$. By looking at the definition of generalized moments~(\ref{eq:moments}) we see that this element is mapped to \[ \langle \hat{a} \rangle = \left\langle f(x_1, x_2 \ldots) \hbar^n \left( \widehat{\Delta x}_1^{n_1} \widehat{\Delta x}_2^{n_2} \ldots \right)_{\rm Weyl} \right\rangle = f(\langle \hat{x}_1\rangle, \langle \hat{x}_2\rangle \ldots) \hbar^n \Delta \left( x_1^{n_1} x_2^{n_2} \ldots \right) \ . \] If $\langle \hat{a} \rangle = 0$, the expectation value is of ``infinite'' order. Otherwise, from definitions~(\ref{eq:orderI})--(\ref{eq:orderIII}) we obtain Order\,$(\langle \hat{a}\rangle ) = 2n + \sum n_i$. Both cases agree with~(\ref{eq:OrderPres}). Since any element of $\mathcal{A}_{ext}$\ can be expressed as a sum of monomial terms of this form, and due to linearity of $\langle.\rangle$\ and properties of Order(.) this extends to all elements of $\mathcal{A}_{ext}$. The extended algebra inherits a non--trivial commutation bracket from the non--abelian associative product of $\mathcal{A}_{kin}$, determined by the CCRs. At the same time, the commutative algebra $\mathcal{A}_{class}$\ can be naturally equipped with the classical Poisson bracket ${\displaystyle \{x_i, x_j\} = \alpha_{ij}^{\ \ k} x_k}\ $. We combine these two brackets to construct an extended formula for computing the quantum Poisson bracket on $\Gamma_Q$\ that is more suitable for directly computing brackets between the generalized moments than equation~(\ref{eq:QPoisson}). We find that for $\hat{f}, \ \hat{g} \in \mathcal{A}_{ext}$ \begin{equation} \label{eq:QPoisson2} \left\{ \langle \hat{f} \rangle,\langle \hat{g} \rangle \right\}_Q = \frac{1}{i\hbar} \left\langle \left[ \hat{f}, \hat{g} \right] \right\rangle + \left\langle \frac{\partial \hat{f}}{\partial x_i} \right\rangle \left\langle \frac{\partial \hat{g}}{\partial x_j} \right\rangle \{x_i, x_j\} + \frac{1}{i\hbar} \left\langle \frac{\partial \hat{f}}{\partial x_i} \right\rangle \left\langle \left[ \hat{x}_i, \hat{g} \right] \right\rangle+ \frac{1}{i\hbar} \left\langle \frac{\partial \hat{g}}{\partial x_i} \right\rangle \left\langle \left[ \hat{f}, \hat{x}_i \right] \right\rangle \ . \end{equation} Here by $\partial \hat{f}/\partial x_i$\ we mean the ordinary partial derivative acting on the elements of $\mathcal{A}_{class}$, so that for a monomial term, with $\hat{a} \in \mathcal{A}_{kin}$\ and $h \in \mathcal{A}_{class}$ \[ \frac{\partial (h\hat{a} )}{\partial x_i} := \frac{\partial h}{\partial x_i} \hat{a} \ . \] It is immediately clear that the formula agrees with the basic definition~(\ref{eq:QPoisson}), when $\hat{f}$\ and $\hat{g}$\ are both in the subalgebra $\mathcal{A}_{kin}$, since in this case only the first term is non-zero. It is also clear upon inspection that the RHS defines a bilinear map over $\mathbb{C}$. Therefore, in order to show that~(\ref{eq:QPoisson2}) holds generally, we only need to show that it holds for a pair of monomials $f\hat{a}$\ and $g \hat{b}$\ with $f, g \in \mathcal{A}_{class}$\ and $\hat{a}, \hat{b} \in \mathcal{A}_{kin}$, since a general element of $\mathcal{A}_{ext}$\ is a linear combination of such terms. Using the quantum Poisson bracket of section~\ref{sec:QPhaseSpace} we have \begin{eqnarray*} \left\{ \langle f\hat{a} \rangle,\langle g\hat{b} \rangle \right\}_Q &=& \left\{ \langle f \rangle \langle \hat{a} \rangle, \langle g \rangle \langle \hat{b} \rangle \right\}_Q \\ &=& \langle f \rangle \langle g \rangle \left\{ \langle \hat{a} \rangle, \langle \hat{b} \rangle \right\}_Q + \langle \hat{a} \rangle \langle \hat{b} \rangle \left\{ \langle f \rangle , \langle g \rangle \right\}_Q + \langle g \rangle \langle \hat{a} \rangle \left\{ \langle f \rangle , \langle \hat{b} \rangle \right\}_Q+ \langle f \rangle \langle \hat{b} \rangle \left\{ \langle \hat{a} \rangle, \langle g \rangle \right\}_Q \\ &=& \langle f \rangle \langle g \rangle \frac{1}{i\hbar} \left\langle \left[ \hat{a}, \hat{b} \right] \right\rangle + \langle \hat{a} \rangle \langle \hat{b} \rangle \left\langle \frac{\partial f}{\partial x_i} \right\rangle \left\langle \frac{\partial g}{\partial x_j} \right\rangle \frac{1}{i\hbar} \left\langle \left[ \hat{x}_i, \hat{x}_j \right] \right\rangle \\ & & + \langle g \rangle \langle \hat{a} \rangle \left\langle \frac{\partial f}{\partial x_i} \right\rangle \frac{1}{i\hbar} \left\langle \left[ \hat{x}_i, \hat{b} \right] \right\rangle + \langle f \rangle \langle \hat{b} \rangle \left\langle \frac{\partial g}{\partial x_i} \right\rangle \frac{1}{i\hbar} \left\langle \left[ \hat{a}, \hat{x}_i \right] \right\rangle \ . \end{eqnarray*} The first equality follows straight from~(\ref{eq:ExpVal}), the second equality utilizing the product rule, which is again used to write the 2nd, 3rd, and 4th terms in the final expression. It is straightforward to see that the right-hand-side of~(\ref{eq:QPoisson2}) yields an identical result. The latter therefore holds for all elements of $\mathcal{A}_{ext}$. We are almost in a position to evaluate the order of the terms on the right-hand side of~(\ref{eq:QPoisson2}) and to verify relation~(\ref{eq:PBorder}) using the properties of ``algebraic'' semiclassical order and the order-preserving property of $\langle . \rangle$. It only remains to understand the effect that the partial derivatives with respect to $x_i$\ have on the semiclassical order. We first note that the derivative does not \emph{reduce} the order of generators $\hbar \hat{\boldsymbol{1}}$\ and $x_i\hat{\boldsymbol{1}}$, while \[ \frac{\partial}{\partial x_i} \widehat{\Delta x}_j = - \delta_{ij} \hat{\boldsymbol{1}} \ , \] so that, acting on any one of these generators of $\mathcal{A}_{ext}$\ , $\partial/\partial x_i$\ reduces the semiclassical order by 1 at the most. In fact, there is an elegant result that is not strictly necessary for our discussion here, but can be quite useful in another context \[ \frac{\partial}{\partial x_i} \left( \widehat{\Delta x}_1^{n_1} \ldots \widehat{\Delta x}_i^{n_i} \ldots \widehat{\Delta x}_M^{n_M} \right)_{\rm Weyl} = - n_i \left( \widehat{\Delta x}_1^{n_1} \ldots \widehat{\Delta x}_i^{n_i-1} \ldots \widehat{\Delta x}_M^{n_M} \right)_{\rm Weyl} \ . \] Since the derivative obeys $\partial (\hat{a} \hat{b})/\partial x_i = (\partial \hat{a} /\partial x_i) \hat{b} + \hat{a} (\partial \hat{b}/\partial x_i )$\ , and using~(\ref{eq:OrderP1}) and~(\ref{eq:OrderP2}), we conclude that the derivative reduces the order of any monomial by 1 at the most. Further, since $\partial/\partial x_i$\ is linear, and again employing~(\ref{eq:OrderP1}) and~(\ref{eq:OrderP2}), we conclude that \begin{equation}\label{eq:OrderP6} {\rm Order\, } \left( \frac{\partial \hat{a} }{\partial x_i} \right) \geq {\rm Order\, } \left( \hat{a} \right) - 1 \ , \quad {\rm for \ any \ \ } \hat{a} \in \mathcal{A}_{ext} \ . \end{equation} Finally, by inspection and referring to properties~(\ref{eq:OrderP1}), ~(\ref{eq:OrderP2})--(\ref{eq:OrderP4}), ~(\ref{eq:OrderP5}), (\ref{eq:OrderPres}), ~(\ref{eq:OrderP6}), we see that the semiclassical order of every term on the right-hand side of~(\ref{eq:QPoisson2}) is bounded from below by ${\rm Order\, } ( \hat{f} ) + {\rm Order\, } ( \hat{g} ) - 2$, which proves~(\ref{eq:PBorder}) for the case where $f = \langle \hat{f} \rangle$\ and $g = \langle \hat{g} \rangle$\ for some $\hat{f}, \hat{g} \in \mathcal{A}_{ext}$\ . Since all of the quantum variables can be written in this way and since they provide a complete coordinate basis on $\Gamma_Q$\ the property~(\ref{eq:PBorder}) \begin{equation*} {\rm Order\, } \left(\{f, g\}_Q \right) \geq {\rm Order\, } \left( f \right) + {\rm Order\, } \left( g \right) -2 \ , \end{equation*} holds for any pair of polynomial functions $f$\ and $g$\ on $\Gamma_Q$\ . For $N\geq2$, this implies that polynomial functions of order $N$\ and above form a \emph{Poisson ideal}, that is, for a polynomial function $f \in \mathscr{F}(\Gamma_Q)$, such that Order $(f) \geq N$ \[ {\rm Order\, } \left( \{f, g\}_Q \right) \geq N \ , \quad {\rm for \ any \ polynomial \ function \ \ } g \in \mathscr{F}(\Gamma_Q) \ . \] The consistency result for the truncated quantum Poisson bracket~(\ref{eq:TruncPBconsistency}) follows immediately for $N\geq1$, since the portions of $f$\ and $g$\ removed by truncation are themselves of order $(N+1)$\ or above and would generate terms of order $(N+1)$\ and above when their quantum Poisson bracket with other functions is taken. \section{Counting truncated constraints} \label{sec:Constraints} In this section we count the truncated set of Casimir constraint functions generated by a single quantum constraint as in~(\ref{eq:CPoly}) order-by-order in the semiclassical expansion. The counting establishes the conditions under which the Casimir constraint removes the number of degrees of freedom equivalent to a single generator. At this stage it is convenient to introduce some further short-hand notation. The Weyl-ordered products of the generators $\widehat{\Delta x}_i$, which form an $\mathcal{A}_{class}$--linear basis on $\mathcal{A}_{ext}$, defined in section~\ref{sec:PBOrder}, will be denoted by \[ \hat{e}_{\vec{i}} = \hat{e}_{(i_1, i_2, \ldots i_M)} := \left(\widehat{\Delta x}_1^{i_1} \widehat{\Delta x}_2^{i_2} \ldots \widehat{\Delta x}_M^{i_M} \right)_{\rm Weyl-ordered} \ . \] Here $\vec{i}$\ is an $M$-tuple of non-negative integers. We will use the convention $\hat{e}_0=\hat{\mathbf{1}}$. For convenience we also define the \textbf{degree} $| \vec{i} | := \sum_{n=1}^M i_n$\ and \textbf{partial ordering} $\vec{i} \geq \vec{j}$\ if $i_n \geq j_n\ \ \forall n$, so that $\vec{i} > \vec{j}$\ if $\vec{i} \geq \vec{j}$\ and $\vec{i} \neq \vec{j}$. We will denote the expectation values of the basis elements by \[ \varepsilon_{(i_1, i_2, \ldots i_M)} = \varepsilon_{\vec{i}} = \langle \hat{e}_{\vec{i}} \rangle \ . \] By construction $\varepsilon_{\vec{i}}=0$, for all $\vec{i}$, with $| \vec{i}| = 1$\ , and $\varepsilon_{0}=1$. We will use $\vec{i}!$\ to denote $(i_1!i_2! \ldots i_M!)$. Since one can symmetrically order an arbitrary product of $\widehat{\Delta x}_i$-s by adding terms that are of lower polynomial degree and proportional to powers of $\hbar$ \begin{eqnarray*} \hat{e}_{\vec{i}} \hat{e}_{\vec{j}} &=& \hat{e}_{(\vec{i}+\vec{j})} + \hbar \sum_{\tiny{\begin{array}{c} \vec{k} < (\vec{i}+\vec{j}) \\ |\vec{k}| \leq |\vec{i}+\vec{j}|-1 \end{array}}} \beta_{\vec{i}, \vec{j}}^{(1) \ \vec{k}} \hat{e}_{\vec{k}} + \hbar^2 \sum_{\tiny{\begin{array}{c} \vec{k} < (\vec{i}+\vec{j}) \\ |\vec{k}| \leq |\vec{i}+\vec{j}|-2 \end{array}}} \beta_{\vec{i}, \vec{j}}^{(2) \ \vec{k}} \hat{e}_{\vec{k}} + \ldots \ , \end{eqnarray*} where $\beta_{\vec{i}, \vec{j}}^{(n) \ \vec{k}}$\ are some polynomials in $x_i$. Starting from section~\ref{sec:counting} we will predominantly work with quantum variables rather than their algebraic analogues, and it will be convenient to use $x_i$\ to denote the expectation values $\langle \hat{x}_i \rangle$\ as well as their ``classical placeholders'' in $\mathcal{A}_{ext}$. In addition, the classical polynomial for the Casimir constraint $C(x_1, x_2, \ldots x_M)$\ will feature prominently and we will denote \[ C := C ( \langle \hat{x}_1 \rangle, \langle \hat{x}_2 \rangle \ldots \langle \hat{x}_M \rangle ) \equiv \langle C(x_1, x_2, \ldots x_M) \hat{\boldsymbol{1}} \rangle \ . \] where $\langle . \rangle$\ in the final expression is taken in the sense of~(\ref{eq:ExpVal}). Similarly we will denote \[ \frac{\partial C}{\partial x_i} := \left\langle \frac{\partial C}{\partial x_i} \hat{\boldsymbol{1}}\right\rangle \ , \quad {\rm and \ so \ forth.} \] \subsection{Constraints and truncation} \label{sec:truncation} As discussed in section~\ref{sec:QLieAlg}, the Casimir constraint can be imposed by demanding that for all polynomial functions $f$, we have $\langle f(\hat{x}_1, \hat{x}_2 \ldots \hat{x}_M) \hat{C} \rangle = 0$. For the purposes of truncation, these conditions can be systematically imposed by using the basis $\{\hat{e}_{\vec{i}} \}$: \[ C_{\vec{i}} := \langle \hat{e}_{\vec{i}\ } \hat{C} \rangle = 0, \ \ \forall \ \vec{i} \in \mathbb{Z}_+^M \ . \] With our conventions $C_0 := \langle \hat{C} \rangle$. This is equivalent to the set of conditions given by equation~(\ref{eq:CPoly}). There are two distinct aspects to the truncation process: 1) truncation of the degrees of freedom; 2) truncation of the above system of the constraint functions. For concreteness, assume we truncate at some order $N \geq 2$\ of the semiclassical expansion. Degrees of freedom are truncated by dropping moments of degree greater than $N$, i.e. drop all $\varepsilon_{\vec{i}}$\ that have $| \vec{i} | > N$. As briefly discussed in section~\ref{sec:Consistent_Trunc}, the truncation of the system of constraints is more subtle. $C_{\vec{i}}$\ are linear functions of the moments $\varepsilon_{\vec{j}}$\ and for the purposes of truncation, we will separate terms appearing in their expressions into three types: \begin{itemize} \item $f(\langle \hat{x}_1 \rangle, \langle \hat{x}_2 \rangle \ldots \langle \hat{x}_M \rangle ) \varepsilon_{\vec{i}}$, where $f$\ is a polynomial in the expectation values, is assigned semiclassical order equal to $| \vec{i} |$. \item $C(\langle \hat{x}_1 \rangle, \langle \hat{x}_2 \rangle \ldots \langle \hat{x}_M \rangle) \varepsilon_{\vec{i}}$, where $C$\ is the classical polynomial expression for the constraint, is assigned, as an exception to the previous point, semiclassical order $\left( | \vec{i} |+ 2 \right)$. \item $\hbar^n f(\langle \hat{x}_1 \rangle, \langle \hat{x}_2 \rangle \ldots \langle \hat{x}_M \rangle) \varepsilon_{\vec{i}}$, arises upon reordering algebra elements and is assigned semiclassical order equal to $\left( | \vec{i} |+ 2n \right)$. \end{itemize} Upon truncation, terms in the expressions for the constraint functions of the semiclassical order higher than $N$\ are dropped, just as prescribed by~(\ref{eq:Trunc1}) and~(\ref{eq:Trunc2}). \subsection{Counting degrees of freedom} \label{sec:counting} First we count the degrees of freedom of an unconstrained system generated by polynomials in $M$\ basic variables. The system is parameterized by $M$\ expectation values and at each semiclassical order $N \geq 2$\ the freedoms are represented by the independent functions $\varepsilon_{\vec{i}}$, where $| \vec{i} | = N$. \emph{How many such variables are there?}---As many as the number of $M$-tuples of non-negative integers with $|\vec{i}|=N$, we will refer to this as $\mathcal{N}_M(N)$. Combinatorially, each such $M$-tuple is produced by considering a row of $N+M-1$\ identical objects and marking $M-1$ of them to serve as partitions, the value of $i_n$\ is then the number of unmarked objects between partition $(n-1)$\ and partition $n$ (where partitions $0$ and $M$\ are assumed to be at the ends). The answer is then simply \[ \mathcal{N}_M(N) = {N+M-1 \choose M-1} \ . \] Imposing the Casimir condition removes a single classical degree of freedom. We thus expect the tower of constraint conditions to remove the equivalent of one combinatorial degree of freedom from the algebra of observables. I.e. after the constraints are imposed, we expect as many degrees of freedom as for a system with $(M-1)$\ generators to remain, so that at each semiclassical order we should then have $\mathcal{N}_{M-1}(N) = {N+M-2 \choose M-2}$\ free variables. Using the identity ${a\choose b} - {a-1 \choose b} = {a-1\choose b-1}$, which is straightforward to verify, we conclude that the required number of independent conditions at each order is ${N+M-2 \choose M-1}$. \emph{How many constraint conditions are there at each order?} This may seem difficult to answer, as constraint conditions $C_{\vec{i}}$\ generally mix terms of different orders. We will proceed by first showing that, when truncated at a given order, the system of constraints becomes finite. The number of constraints at each order is then the number of additional non-trivial constraint conditions that arise when we raise the truncation order by one. Analogously to relation~(\ref{eq:monomial_exp}), by writing $\hat{x}_i = \widehat{\Delta x}_i + x_i\mathbf{\hat{1}}$, we can expand any symmetrized monomial as \begin{equation}\label{eq:C_expansion} \left( \hat{x}_1^{i_1} \hat{x}_2^{i_2} \ldots \right)_{\rm Weyl} = \sum_{\vec{j} \leq \vec{i}} \frac{1}{\vec{j}!} \frac{\partial^{|\vec{j}|} \left( x_1^{i_1} x_2^{i_2} \ldots \right)}{\partial^{j_1} x_1 \partial^{j_2} x_2 \ldots } \hat{e}_{\vec{j}} \ . \end{equation} This easily extends to polynomials that are sums of symmetrized monomials. Of course, any monomial can be symmetrized by adding lower polynomial order terms proportional to powers of $\hbar$. In particular, the constraint operator itself can be expressed in the form \[ \hat{C} = \sum_{\vec{j} \leq \vec{i}} \frac{1}{\vec{j}!} \frac{\partial^{|\vec{j}|} C}{\partial^{j_1} x_1 \partial^{j_2} x_2 \ldots } \hat{e}_{\vec{j}} + \hbar \times \left( {\rm terms\ of\ order\ 0\ and\ higher} \right) \ . \] A general constraint function therefore has the form \begin{eqnarray}\label{eq:constraints} C_{\vec{i}} &=& \sum_{\vec{j} \geq \vec{i}} \frac{1}{(\vec{j} - \vec{i})!} \frac{\partial^{|\vec{j} - \vec{i}|} C}{\partial x_1^{j_1-i_1} \ldots \partial x_{M}^{j_M-i_M}} \varepsilon_{\vec{j}} + \hbar \sum_{|\vec{j}| \geq |\vec{i}|-1} \alpha_{C_{\vec{i}}}^{(1) \ \vec{j}} \varepsilon_{\vec{j}} \nonumber \\ & & + \hbar^2 \sum_{|\vec{j}| \geq |\vec{i}|-2} \alpha_{C_{\vec{i}}}^{(2) \ \vec{j}} \varepsilon_{\vec{j}} + \ldots + \hbar^{|\vec{i}|} \sum_{\vec{j}} \alpha_{C_{\vec{i}}}^{(|\vec{i}|) \ \vec{j}} \varepsilon_{\vec{j}} \ . \end{eqnarray} Here $\alpha_{C_{\vec{i}}}^{(n) \ {\vec{j}}}$\ are coefficients polynomial in the expectation values $x_i$, and thus of semiclassical order $0$ (not to be confused with the Lie algebra structure constants $\alpha_{ij}^{\ k}$\ of sections~\ref{sec:QLieAlg} and~\ref{sec:PBOrder}). The first sum comes from the Weyl-symmetric part of the element $\hat{e}_{\vec{i}\ } \hat{C}$, subsequent sums arise from its components that are antisymmetric in one, two and more adjacent pairs of moment-generating elements $\widehat{\Delta x}_i$: each antisymmetric pair can be reduced by using the CCRs thus producing the powers of $\hbar$. Additional terms in the second, third and further sums come from the expectation value of the product between $\hat{e}_{\vec{i}}$\ and terms multiplied by $\hbar$\ in the expression~(\ref{eq:C_expansion}) for the constraint element: such terms are of order $(|\vec{i}|+2)$\ or higher. The important feature of the above expansion is that the lowest semiclassical order terms in the first sum are $C\varepsilon_{\vec{i}}$\ and $\sum_k \frac{\partial C}{\partial x_k} \varepsilon_{(i_1, \ldots, i_k+1, \ldots, i_M)}$. Since for the purposes of truncating the constraints $C$\ is of order $2$, the latter term has the lowest order, which is $(|\vec{i}|+1)$. Terms coming from the second sum are at least of the order $(|\vec{i}|+1)$, the rest of the sums contribute terms of order $(|\vec{i}|+2)$\ and above. Thus, after truncation at order $N$, constraints $C_{\vec{i}}=0$\ are satisfied identically for all $|\vec{i}| >N-1$. The number of non-trivial conditions up to order $N$\ is the same as the number of non-negative integer $M$-tuples of degree $N-1$\ and less. The change in this number as we go from truncation at order $N-1$\ to truncation at order $N$\ is the same as the number of $M$-tuples of degree $N-1$, namely $\mathcal{N}_M(N-1)={N+M-2 \choose M-1}$\ as required by the counting. This shows that, provided the non-trivial constraint conditions remaining after truncation are functionally independent, they remove precisely one combinatorial degree of freedom. In the next section, we prove that under a reasonable set of conditions, which is sufficiently broad for our purposes, the truncated set of constraints is indeed functionally independent. \subsection{Independence of truncated constraints} In this section we look for conditions under which the set of constraint functions $\{C_{\vec{i}}\}$\ truncated at some order $N$\ in accordance with Section~\ref{sec:truncation} is \emph{functionally independent}. We first formulate the conditions needed for the correct reduction of degrees of freedom. Then, using semiclassical orders, we argue that a simpler set of conditions, more amenable to direct analysis, is sufficient for semiclassical states. Detailed analysis is carried out in subsection~\ref{sec:theorems}.\\ In general, we consider a (finite) set of functions $\{f_i\}_{i=1, \ldots L}$\ \emph{functionally} independent in some region if fixing the values of all these functions defines a hypersurface of codimension $L$. That is, locally these functions fix $L$\ degrees of freedom---precisely the desired property for our system of constraints. We can apply a version of the Frobenius Integrability Theorem to the set of exterior derivatives $\{ df_i \}_{i=1, \ldots L}$: the functions integrate to a hypersurface of codimension $L$\ if and only if the set of their exterior derivatives is \emph{linearly} independent at every point in the region of interest. We assume that in the original classical system the single (classical) constraint removes a single classical degree of freedom. By the integrability considerations, this implies that the constraint polynomial must be regular on the constraint surface, i.e. $dC|_{\tiny{C=0}} \neq 0$. By continuity, this must also hold in some neighborhood of $C=0$. Thus, for correct reduction of the degrees of freedom, we need to show that $\{d_QC_{\vec{j}} \}_{|\vec{i}|\leq N-1}$\ are linearly independent. Here we treat the quantum phase space $\Gamma_Q$, prior to imposing the constraint, as the cartesian product $\Gamma_Q=\Gamma \times \mathcal{M}_{\varepsilon}$ of the classical phase space $\Gamma$\ and the space of moments $\mathcal{M}_{\varepsilon}$. The exterior derivative on $\Gamma_Q$\ is the (direct) sum of exterior derivatives on the component spaces $d_Q = d + d_{\varepsilon}$. The covariant coordinate vectors are $d_Qx_i = dx_i$\ and $d_Q\varepsilon_{\vec{i}} = d_{\varepsilon} \varepsilon_{\vec{i}}$. Naturally a pair of non-zero gradients $df$\ and $d_{\varepsilon}g$\ are always linearly independent as they belong to different disjoint subspaces of the cotangent space. In analyzing semiclassical orders below we treat all non-truncated coordinate directions on equal footing. Thus the statement ``$d_Qf$\ has leading contribution of order $n$'' will mean that the \emph{coefficient} for one of the coordinate covectors $dx_i$\ or $d_{\varepsilon} \varepsilon_{\vec{i}}$\ in the coordinate decomposition of $d_Qf$\ is of semiclassical order $n$, while others have coefficients of equal or higher order. Let us specialize the expression for the general constraint function $C_{\vec{i}}$~(\ref{eq:constraints}) to truncation at order $N$\ as described in Section~\ref{sec:truncation}. For $|\vec{i}| < N-1$\ the sums over $\vec{j}$\ terminate when $|\vec{j}|=N-2n$, where $n$\ is the power of $\hbar$, multiplying the sum. In particular, \begin{eqnarray}\label{eq:trnc_constraints} C_{\vec{i}} &=& \sum_{\tiny{\begin{array}{c} \vec{j} \geq \vec{i} \\ 1< |\vec{j}| \leq N \end{array}}} \frac{1}{(\vec{j} - \vec{i})!} \frac{\partial^{|\vec{j} - \vec{i}|} C}{\partial x_1^{j_1-i_1} \ldots \partial x_{M}^{j_M-i_M}} \varepsilon_{\vec{j}} \ + \ \hbar \sum_{N-2\geq |\vec{j}| \geq |\vec{i}|-1} \alpha_{C_{\vec{i}}}^{(1) \ \vec{j}} \varepsilon_{\vec{j}} \nn\\ && + \ \ \hbar^2\,\, \sum_{N-4 \geq |\vec{j}| \geq |\vec{i}|-2} \alpha_{C_{\vec{i}}}^{(2) \ \vec{j}} \varepsilon_{\vec{j}} + \ldots \ . \end{eqnarray} For $|\vec{i}| = N-1$\ the sums terminate in the same way, however the first term in the expansion $C\varepsilon_{\vec{i}}$\ is treated as having semiclassical order $|\vec{i}|+2 = N+1 >N$\ and thus is dropped upon truncation. \begin{eqnarray}\label{eq:trnc_constraints2} C_{\vec{i}} &=& \sum_{\tiny{\begin{array}{c} \vec{j} > \vec{i} \\ 1< |\vec{j}| \leq N \end{array}}} \frac{1}{(\vec{j} - \vec{i})!} \frac{\partial^{|\vec{j} - \vec{i}|} C}{\partial x_1^{j_1-i_1} \ldots \partial x_{M}^{j_M-i_M}} \varepsilon_{\vec{j}} + \hbar \sum_{|\vec{j}|=N-2} \alpha_{C_{\vec{i}}}^{(1) \ \vec{j}} \varepsilon_{\vec{j}} \nn\\ &=& \sum_{l=1}^{M} \frac{\partial C}{\partial x_l} \varepsilon_{(i_1, \ldots, i_l+1, \ldots)} + \hbar \sum_{|\vec{j}|=N-2} \alpha_{C_{\vec{i}}}^{(1) \ \vec{j}} \varepsilon_{\vec{j}} \ . \end{eqnarray} Higher reordering terms get dropped as they are of order higher than $N$. From~(\ref{eq:trnc_constraints}) and~(\ref{eq:trnc_constraints2}) we infer the following qualitative features of $d_QC_{\vec{i}} = dC_{\vec{i}} + d_{\varepsilon}C_{\vec{i}}$: \begin{itemize} \item leading order contribution to $dC_0$\ comes from $dC$\ and is of semiclassical order $0$ \item leading order contribution to all other $dC_{\vec{i}}$\ is at least of semiclassical order $2$ \item contributions to $d_{\varepsilon} C_{\vec{i}}$\ start at order $0$. \end{itemize} We infer, that to leading semiclassical order, $d_QC_0$\ is linearly independent of $\{ d_Q C_{\vec{i}}\}_{1\leq|\vec{i}|\leq N-1}$. Furthermore, to establish linear independence of $\{d_Q C_{\vec{i}}\}_{1\leq|\vec{i}|\leq N-1}$\ to leading semiclassical order we only need to concern ourselves with the linear independence of $\{ d_{\varepsilon}C_{\vec{i}} \}_{1\leq |\vec{i}|\leq N-1}$. Taking gradient of the truncated constraint functions with respect to the moments we obtain, for $|\vec{i}| < N-1$: \begin{eqnarray*} d_{\varepsilon}C_{\vec{i}} &=& \sum_{\tiny{\begin{array}{c} \vec{j} \geq \vec{i} \\ 1<|\vec{j}| \leq N \end{array}}} \frac{1}{(\vec{j} - \vec{i})!} \frac{\partial^{|\vec{j} - \vec{i}|} C}{\partial x_1^{j_1-i_1} \ldots \partial x_{M}^{j_M-i_M}} d_{\varepsilon} \varepsilon_{\vec{j}} + \hbar \sum_{N-2 \geq |\vec{j}| \geq |\vec{i}|-1} \alpha_{C_{\vec{i}}}^{(1) \ \vec{j}} d_{\varepsilon} \varepsilon_{\vec{j}}\\ &&+ \hbar^2 \sum_{N-4 \geq |\vec{j}| \geq |\vec{i}|-2} \alpha_{C_{\vec{i}}}^{(2) \ \vec{j}} d_{\varepsilon} \varepsilon_{\vec{j}} + \ldots \ . \end{eqnarray*} A similar expression follows for $|\vec{i}| = N-1$\ by using~(\ref{eq:trnc_constraints2}). In either case, the leading order contribution to the gradient along each direction $d_{\varepsilon} \varepsilon_{\vec{i}}$, is given by the terms in the first sum alone, other contributions are suppressed by powers of $\hbar$. We will denote this `symmetric' part of $C_{\vec{i}}$\ by $\tilde{C}_{\vec{i}}$\ so that for $|\vec{i}| < N-1$: \begin{eqnarray*} d_{\varepsilon}\tilde{C}_{\vec{i}} = \sum_{\tiny{\begin{array}{c} \vec{j} \geq \vec{i} \\ 1 < |\vec{j}| \leq N \end{array}}} \frac{1}{(\vec{j} - \vec{i})!} \frac{\partial^{|\vec{j} - \vec{i}|} C}{\partial x_1^{j_1-i_1} \ldots \partial x_{M}^{j_M-i_M}} d_{\varepsilon} \varepsilon_{\vec{j}} \ . \end{eqnarray*} A similar expression follows for $|\vec{i}| = N-1$. As all the constraint functions are linear in the moments $\varepsilon_{\vec{i}}$, the gradients $d_{\varepsilon}C_{\vec{i}}$\ as well as their symmetric parts $d_{\varepsilon} \tilde{C}_{\vec{i}}$\ have coefficients that depend on the expectation values alone, and are thus much easier to analyze than $d_QC_{\vec{i}}$.\\ In the next subsection we focus on linear independence of $\{ d_{\varepsilon}\tilde{C}_{\vec{i}} \}_{1\leq |\vec{i}|\leq N-1}$\ and briefly touch upon the way higher order corrections may affect the results. From the results established below we conclude, that for a sufficiently semiclassical state: \begin{itemize}\item for expectation values satisfying the classical constraint function, the truncated quantum constraints are functionally independent so long as $dC$\ is not comparable to $\hbar$\ or the moments in at least one coordinate direction \item for expectation values off the classical constraint surface, the constraint functions are functionally independent so long as for some $k$\ neither $\frac{\partial C}{\partial x_k}$\ nor $\frac{\partial^{N-2}}{\partial x_k^{N-2}}\left( \frac{1}{C} \right)$\ are comparable to $\hbar$\ or the moments. \end{itemize} While these conditions can be violated, this is likely to happen sufficiently ``far away'' from the classical constraint surface as near $C=0$, $\frac{1}{C}$\ and its derivatives blow up. In addition, these conditions are \emph{sufficient}, but \emph{not necessary} and in some cases the constraints may be independent even if they do not hold. In any event, the above conditions provide a viable and rigid test of whether the semiclassical truncation of a given constrained system reduces the degrees of freedom correctly. \subsection{Details of the argument} \label{sec:theorems} \textbf{Key result}: \emph{for the system of constraints truncated at order $N$, the set of gradients $\{ d_{\varepsilon}\tilde{C}_{\vec{i}} \}_{1\leq |\vec{i}|\leq N-1}$\ is linearly independent when expectation values $x_i$\ lie in some neighborhood of the classical constraint surface $C=0$.} (Note: this does not place any conditions on the values of moments.)\\ \ \\ \textbf{Proof}: We need only prove that the gradients are linearly independent when the expectation values \emph{do} lie on the classical constraint surface. By continuity, they remain independent in some open neighborhood of the surface $C=0$.\\ \ \\ Suppose that the set $\{ d_{\varepsilon}\tilde{C}_{\vec{i}} \}_{1\leq |\vec{i}|\leq N-1}$\ is linearly dependent at some point $P\in \Gamma_Q$, then there are numerical coefficients (with appropriate units) $\gamma^{\vec{i}}$\ such that \[ \sum_{1\leq|\vec{i}|\leq N-1} \left. \gamma^{\vec{i}} d_{\varepsilon} \tilde{C}_{\vec{i}} \right|_{\tiny{P}} = 0 \ . \] However, \begin{eqnarray*} \sum_{1\leq|\vec{i}|\leq N-1} \left. \gamma^{\vec{i}}d_{\varepsilon}\tilde{C}_{\vec{i}} \right|_{\tiny{P}} &=& \sum_{1\leq|\vec{i}| < N-1} \sum_{\tiny{\begin{array}{c} \vec{j} \geq \vec{i}\\ 1<|\vec{j}| \leq N \end{array}}} \left. \gamma^{\vec{i}} \frac{1}{(\vec{j} - \vec{i})!} \frac{\partial^{|\vec{j} - \vec{i}|} C}{\partial x_1^{j_1-i_1} \ldots \partial x_{M}^{j_M-i_M}} d_{\varepsilon} \varepsilon_{\vec{j}} \right|_{\tiny{P}} \\ && + \sum_{|\vec{i}| = N-1} \sum_{\tiny{\begin{array}{c} \vec{j} > \vec{i}\\ |\vec{j}| = N \end{array}}} \left. \gamma^{\vec{i}} \frac{1}{(\vec{j} - \vec{i})!} \frac{\partial^{|\vec{j} - \vec{i}|} C}{\partial x_1^{j_1-i_1} \ldots \partial x_{M}^{j_M-i_M}} d_{\varepsilon} \varepsilon_{\vec{j}} \right|_{\tiny{P}} \ . \end{eqnarray*} Two separate (double) sums appear due to the subtlety of truncation, since the terms $C\varepsilon_{\vec{j}}$\ are dropped for $|\vec{j}|\geq N-1$, so that the terms with $\vec{j}=\vec{i}$\ do not appear for $|\vec{i}|=N-1$. Notice that in the first (double) sum above $|\vec{i}| < N-1$\ so that, when $|\vec{j}|=N-1$\ or $|\vec{j}|=N$, we always have $\vec{j}>\vec{i}$. We combine the $|\vec{j}|=N-1$\ and $|\vec{j}|=N$\ terms from the first double sum with the second double sum to obtain \begin{eqnarray*} \sum_{1\leq|\vec{i}|\leq N-1} \left. \gamma^{\vec{i}}d_{\varepsilon}\tilde{C}_{\vec{i}} \right|_{\tiny{P}} &=& \sum_{1\leq|\vec{i}| < N-1} \sum_{\tiny{\begin{array}{c} \vec{j} \geq \vec{i}\\ 1<|\vec{j}| < N-1 \end{array}}} \left. \gamma^{\vec{i}} \frac{1}{(\vec{j} - \vec{i})!} \frac{\partial^{|\vec{j} - \vec{i}|} C}{\partial x_1^{j_1-i_1} \ldots \partial x_{M}^{j_M-i_M}} d_{\varepsilon} \varepsilon_{\vec{j}} \right|_{\tiny{P}} \\ && + \sum_{1 \leq |\vec{i}| \leq N-1} \sum_{\tiny{\begin{array}{c} \vec{j} > \vec{i}\\ N-1 \leq |\vec{j}| \leq N \end{array}}} \left. \gamma^{\vec{i}} \frac{1}{(\vec{j} - \vec{i})!} \frac{\partial^{|\vec{j} - \vec{i}|} C}{\partial x_1^{j_1-i_1} \ldots \partial x_{M}^{j_M-i_M}} d_{\varepsilon} \varepsilon_{\vec{j}} \right|_{\tiny{P}} \ . \end{eqnarray*} We rewrite this expression by reversing the order of summation: summing first over $\vec{j}$\ then over $\vec{i}$. \begin{eqnarray*} \sum_{1\leq|\vec{i}|\leq N-1} \left. \gamma^{\vec{i}}d_{\varepsilon}\tilde{C}_{\vec{i}} \right|_{\tiny{P}} &=& \sum_{1<|\vec{j}| < N-1} \left. \left( \sum_{0<\vec{i} \leq \vec{j}} \gamma^{\vec{i}} \frac{1}{(\vec{j} - \vec{i})!} \frac{\partial^{|\vec{j} - \vec{i}|} C}{\partial x_1^{j_1-i_1} \ldots \partial x_{M}^{j_M-i_M}} \right) d_{\varepsilon} \varepsilon_{\vec{j}} \right|_{\tiny{P}}\\ && + \sum_{N-1 \leq |\vec{j}| \leq N} \left. \left( \sum_{0<\vec{i} < \vec{j}} \gamma^{\vec{i}} \frac{1}{(\vec{j} - \vec{i})!} \frac{\partial^{|\vec{j} - \vec{i}|} C}{\partial x_1^{j_1-i_1} \ldots \partial x_{M}^{j_M-i_M}} \right) d_{\varepsilon} \varepsilon_{\vec{j}} \right|_{\tiny{P}} \ . \end{eqnarray*} We can now clearly read off the coefficient in front of each coordinate gradient $d_{\varepsilon} \varepsilon_{\vec{j}}$\ in the sum. Since each $d_{\varepsilon} \varepsilon_{\vec{j}}$\ is an independent covariant coordinate vector, the coefficients must vanish independently. For each $\vec{j}$\ satisfying $1<|\vec{j}| < N-1$\ we then have a condition: \begin{equation}\label{eq:coeffs1} \sum_{ 0< \vec{i} \leq \vec{j}} \left. \gamma^{\vec{i}} \frac{1}{(\vec{j} - \vec{i})!} \frac{\partial^{|\vec{j} - \vec{i}|} C}{\partial x_1^{j_1-i_1} \ldots \partial x_{M}^{j_M-i_M}} \right|_{\tiny{P}} = 0 \ , \end{equation} while for $N-1\leq |\vec{j}| \leq N$\ we have \begin{equation}\label{eq:coeffs2} \sum_{ 0< \vec{i} < \vec{j}} \left. \gamma^{\vec{i}} \frac{1}{(\vec{j} - \vec{i})!} \frac{\partial^{|\vec{j} - \vec{i}|} C}{\partial x_1^{j_1-i_1} \ldots \partial x_{M}^{j_M-i_M}} \right|_{\tiny{P}} = 0 \ . \end{equation} Now we assume that $P$\ lies on the classical constraint surface and that the classical constraint is regular, i.e. $\left. C \right|_{\tiny{P}} = 0$\ and $\left. dC \right|_{\tiny{P}} \neq 0$. As an immediate consequence, terms proportional to $C$\ drop out in the sum~(\ref{eq:coeffs1}) and thus one can use condition~(\ref{eq:coeffs2}) for all $1<|\vec{j}|\leq N$. We proceed in two steps: 1) show that~(\ref{eq:coeffs2}) implies that all $\gamma^{\vec{i}}=0$\ for $|\vec{i}|=1$; 2) use induction to conclude that $\gamma^{\vec{i}}=0$\ for all $\vec{i}$. \\ \ \\ \emph{Step 1}. Since $\left. dC \right|_{\tiny{P}} \neq 0$, there is a direction along which the derivative of $C$\ does not vanish at $P$, i.e. for some $k\in \{1, 2, \ldots, M\}$, we have $\left. \frac{\partial C}{\partial x_k} \right|_{\tiny{P}} \neq 0$. Denote the coefficients $\gamma_{(k)}^m:= \gamma^{\vec{i}}$, where all but the $k$-th entry in the $M$-tuple $\vec{i}$\ are zero, namely $i_n = m \delta_{nk}$. For these coefficients, with $m \leq N$\ the condition~(\ref{eq:coeffs2}) takes form \begin{equation}\label{eq:condition2_spec} \sum_{n=1}^{m-1} \left. \gamma_{(k)}^n \frac{1}{(m-n)!} \frac{\partial^{m-n}C}{\partial x_k^{m-n}} \right|_{\tiny{P}} = 0 \ . \end{equation} From our assumptions and the above relation evaluated at $m=2$\ we immediately conclude that $\gamma_{(k)}^1=0$. This is true for any $k$\ such that $\left. \frac{\partial C}{\partial x_k} \right|_{\tiny{P}} \neq 0$. Now consider any $l\neq k$. Denote the coefficients $\gamma_{(k, l)}^{r\ s} = \gamma^{\vec{i}}$, where $k$-th and $l$-th entries of the $M$-tuple $\vec{i}$\ are $r$\ and $s$\ respectively, while the rest are zero, i.e. $i_n = r\delta_{kn} + s\delta_{ln}$. (Note that $\gamma_{(k, l)}^{r\ 0} = \gamma_{(k)}^r$). Let $\vec{j}$\ have components $j_n = \delta_{kn} + \delta_{ln}$, the condition~(\ref{eq:coeffs2}) becomes \[ \left. \gamma_{(k, l)}^{1\ 0} \frac{\partial C}{\partial x_l} \right|_{\tiny{P}} + \left. \gamma_{(k, l)}^{0\ 1} \frac{\partial C}{\partial x_k} \right|_{\tiny{P}} = 0 \ . \] The first term vanishes as $\gamma_{(k, l)}^{1\ 0} = \gamma_{(k)}^1=0$, while $\left. \frac{\partial C}{\partial x_k} \right|_{\tiny{P}} \neq 0$\ and it follows that $\gamma_{(k, l)}^{0\ 1}=0$. Thus for all $\vec{i}$\ satisfying $|\vec{i}|=1$\ condition~(\ref{eq:coeffs2}) together with $C|_{\tiny{P}}=0$\ imply $\gamma^{\vec{i}}=0$.\\ \ \\ \emph{Step 2}. We now assume that $\gamma^{\vec{i}}=0$\ for all $|\vec{i}| \leq m$. Evaluating the sum~(\ref{eq:coeffs2}) for some $\vec{j}$\ such that $|\vec{j}|=m+2$, and dropping the terms multiplied by $\gamma^{\vec{i}}$\ with $|\vec{i}| \leq m$\ we obtain \begin{equation}\label{eq:n_plus_2} \sum_{\tiny{\begin{array}{c} l=1\\ j_l \neq 0 \end{array}}}^M \left. \gamma^{\vec{j} - \vec{v}_{(l)}} \frac{\partial C}{\partial x_l} \right|_{\tiny{P}} = 0 \ , \end{equation} where $\vec{v}_{(l)}$\ is the $M$-tuple with a single non-zero unit entry $\vec{v}_{(l)\ n} = \delta_{ln}$. Setting $\vec{j} = (m+2)\vec{v}_{(k)}$, where $k$, as before, labels a direction along which $\left. \frac{\partial C}{\partial x_k} \right|_{\tiny{P}} \neq 0$, the sum~(\ref{eq:n_plus_2}) reduces to a single term \[ \left. \gamma^{(m+1)\vec{v}_{(k)}} \frac{\partial C}{\partial x_k} \right|_{\tiny{P}} = 0 \quad {\rm implying} \quad \gamma^{(m+1)\vec{v}_{(k)}}=0 \ . \] Consider the following `step' that takes us from $(m+1)\vec{v}_{(k)}$\ to another $M$-tuple of degree $(m+1)$: subtract $\vec{v}_{(k)}$\ and add $\vec{v}_{(l)}$, for some $l\neq k$. Naturally, the first step yields $m\vec{v}_{(k)} + \vec{v}_{(l)}$\ and we note that any $M$-tuple of degree $(m+1)$ can be reached by starting at $(m+1)\vec{v}_{(k)}$\ and taking up to $(m+1)$\ such steps. Using $\gamma^{(m+1)\vec{v}_{(k)}}=0$, it is straightforward to verify that $\gamma^{\vec{i}}=0$\ for all $\vec{i}$\ that are one step away from $(m+1)\vec{v}_{(k)}$, we only need to set $\vec{j} = (m+1)\vec{v}_{(k)} + \vec{v}_{(l)}$, and use~(\ref{eq:n_plus_2}) to get \[ \left. \gamma^{m\vec{v}_{(k)}+\vec{v}_{(l)}} \frac{\partial C}{\partial x_k} \right|_{\tiny{P}} + \left. \gamma^{(m+1)\vec{v}_{(k)}} \frac{\partial C}{\partial x_l} \right|_{\tiny{P}} = 0 \ . \] Since $\gamma^{(m+1)\vec{v}_{(k)}}=0$, this immediately yields $ \gamma^{m\vec{v}_{(k)}+\vec{v}_{(l)}} = 0$. The process can be continued iteratively: we notice that $\vec{i}$\ being $r$\ steps away from $(m+1)\vec{v}_{(k)}$\ implies $i_k = (m+1-r)$. Suppose that $\gamma^{\vec{i}}=0$\ for all $\vec{i}$\ that are up to $r$\ steps away from $(m+1)\vec{v}_{(k)}$. Pick some $\vec{i}$\ that is exactly $r$\ steps away from $(m+1)\vec{v}_{(k)}$, so that $i_k = (m+1-r)$, and evaluate~(\ref{eq:n_plus_2}) with $\vec{j} = (\vec{i} + \vec{v}_{(l)})$\ for any $l\neq k$ \[ \sum_{\tiny{\begin{array}{c} l'=1\\ i_{l'}+\delta_{ll'} \neq 0 \end{array}}} \left. \gamma^{\vec{i} + \vec{v}_{(l)} - \vec{v}_{(l')}} \frac{\partial C}{\partial x_l} \right|_{\tiny{P}} = 0 \ . \] Since $l\neq k$, $(\vec{i} + \vec{v}_{(l)} - \vec{v}_{(l')})_k = (m+1 - (r+\delta_{l'k}))$, so that all but one `$\gamma$' are exactly $r$\ steps away from $(m+1)\vec{v}_{(k)}$, and hence vanish, leaving us with \[ \left. \gamma^{\vec{i} - \vec{v}_{(k)}+ \vec{v}_{(l)}} \frac{\partial C}{\partial x_k} \right|_{\tiny{P}} = 0 \quad {\rm implying} \quad \gamma^{\vec{i} - \vec{v}_{(k)}+ \vec{v}_{(l)}}=0 \ . \] From here it quickly follows that $\gamma^{\vec{i}}=0$\ for all $\vec{i}$\ that are up to $(r+1)$\ steps away from $(m+1)\vec{v}_{(k)}$. Thus by induction it first follows that $\gamma^{\vec{i}}=0$\ for all $|\vec{i}|=m+1$\ and therefore, again inductively, for all $|\vec{i}|\leq N$. This completes the proof.\\ \ \\ \textbf{Corollary:} \emph{for the system of constraints truncated at orders $N=2$\ or $N=3$, the set of gradients $\{ d_{\varepsilon}\tilde{C}_{\vec{i}} \}_{1\leq |\vec{i}|\leq N-1}$\ is linearly independent everywhere, where $C$\ is regular.}\\ \ \\ This follows as at these orders, no terms proportional to $C$\ appear in the constraints $\{ C_{\vec{i}} \}_{1\leq |\vec{i}|\leq N-1}$\ and linear dependence leads to~(\ref{eq:coeffs2})\ for all $\vec{j}$\ hence the above proof applies regardless of the value taken by $C$.\\ \ \\ While this result guarantees that there is some interesting range of values for which the constraint functions are functionally independent for sufficiently semiclassical states, it does not provide a method for verifying whether the gradients are linearly independent for a given set of values off the constraint surface taken by $x_i$. Fortunately, the result can be made somewhat stronger.\\ \ \\ \textbf{Stronger statement}: \emph{for the system of constraints truncated at order $N$ the set of gradients $\{ d_{\varepsilon}\tilde{C}_{\vec{i}} \}_{1\leq |\vec{i}|\leq N-1}$\ is linearly dependent at some point $P$, with $C|_{\tiny{P}} \neq 0$, only if $\left. \frac{\partial^{N-2} }{\partial x_k^{N-2}}\left(\frac{1}{C} \right) \right|_{\tiny{P}} = 0$\ for every $k$\ such that $\left. \frac{\partial C}{\partial x_k} \right|_{\tiny{P}} \neq 0$.}\\ \ \\ \textbf{Proof}: we show that the above condition is necessary for \emph{linear dependence} of the gradients $\{ d_{\varepsilon}\tilde{C}_{\vec{i}} \}_{1\leq |\vec{i}|\leq N-1}$\ by demonstrating that its converse implies \emph{linear independence}.\\ \ \\ Specifically, let $P$\ be such that $C|_{\tiny{P}}\neq 0$\ and suppose there is some direction labeled $k$\ such that $\left. \frac{\partial C}{\partial x_k} \right|_{\tiny{P}} \neq 0$\ as well as $\left. \frac{\partial^{N-2} }{\partial x_k^{N-2}}\left(\frac{1}{C} \right) \right|_{\tiny{P}} \neq 0$. Suppose, as in the previous proof, that there are some coefficients $\gamma^{\vec{i}}$\ such that \[ \sum_{1\leq|\vec{i}|\leq N-1} \left. \gamma^{\vec{i}} d_{\varepsilon} \tilde{C}_{\vec{i}} \right|_{\tiny{P}} = 0 \ . \] We repeat the steps from the previous proof leading up to the relation~(\ref{eq:condition2_spec}). This time, since $C|_{\tiny{P}} \neq 0$\ we need to consider both types of conditions~(\ref{eq:coeffs1}) and~(\ref{eq:coeffs2}). The former condition takes the form of a recursion relation for $m<N-1$ \begin{equation}\label{eq:condition1_spec} \left. \gamma_{(k)}^m C \right|_{\tiny{P}} = - \sum_{n=1}^{m-1} \left. \gamma_{(k)}^n \frac{1}{(m-n)!} \frac{\partial^{m-n}C}{\partial x_k^{m-n}} \right|_{\tiny{P}} \ . \end{equation} For $m=N-1$\ and $m=N$\ relation~(\ref{eq:condition2_spec}) still holds, since terms proportional to $C$\ do not appear in~(\ref{eq:coeffs2}). Using $m=N-1$\ in~(\ref{eq:condition2_spec}) we obtain \begin{equation}\label{eq:condition1_Nmin1} \sum_{n=1}^{N-2} \left. \gamma_{(k)}^n \frac{1}{(N-n-1)!} \frac{\partial^{N-n-1}C}{\partial x_k^{N-n-1}} \right|_{\tiny{P}} = 0 \ . \end{equation} As we prove in the appendix, (\ref{eq:condition1_spec}) implies that for $m<N-1$ \begin{equation} \label{eq:recursion_result} \gamma_{(k)}^m = \left. \frac{\gamma_{(k)}^1 C}{(m-1)!} \frac{\partial^{m-1}}{\partial x_k^{m-1}}\left( \frac{1}{C} \right) \right|_{\tiny{P}} \ , \end{equation} which, substituted into~(\ref{eq:condition1_Nmin1}), immediately gives \[ 0 = \left. \frac{\gamma_{(k)}^1}{(N-2)!} \frac{\partial^{N-2}}{\partial x_k^{N-2}}\left( \frac{1}{C} \right) \right|_{\tiny{P}} \ . \] Therefore, either $\gamma_{(k)}^1 =0$\ or $\left.\frac{\partial^{N-2}}{\partial x_k^{N-2}} \left( \frac{1}{C} \right) \right|_{\tiny{P}} = 0$\ , and by our assumption we conclude that $\gamma_{(k)}^1 =0$. Immediately, from~(\ref{eq:recursion_result}) we conclude that $\gamma_{(k)}^m=0$\ for $1 \leq m <N-1$. Finally, we use the relation~(\ref{eq:condition2_spec}) setting $m=N$\ to obtain \[ \gamma_{(k)}^{N-1} \left. \frac{\partial C}{\partial x_k} \right|_{\tiny{P}} = -\sum_{n=1}^{N-2} \left.\gamma_{(k)}^n \frac{1}{(N-n)!} \frac{\partial^{N-n}C}{\partial x_k^{N-n}} \right|_{\tiny{P}} \ . \] The right-hand side vanishes as all the terms are proportional to $\gamma_{(k)}^{m}$\ with $1\leq m <N-1$, and hence $\gamma_{(k)}^{N-1}=0$, which now accounts for all the coefficients $\gamma_{(k)}^{m}$. Let us now consider some $l \neq k$\ and specialize condition~(\ref{eq:coeffs1}) to $\vec{j}$\ with $j_n = m\delta_{kn} + \delta_{ln}$. With the same notation as in the previous proof, we obtain for $1<m+1<N-1$ \[ \sum_{n=1}^m \left. \gamma_{(k,l)}^{n\ 0} \frac{1}{(m-n)!} \frac{\partial^{m-n+1} C}{\partial x_k^{m-n} \partial x_l} \right|_{\tiny{P}} + \sum_{n=0}^m \left. \gamma_{(k,l)}^{n\ 1} \frac{1}{(m-n)!} \frac{\partial^{m-n} C}{\partial x_k^{m-n}} \right|_{\tiny{P}} =0 \ . \] The entire first sum vanishes since $\gamma_{(k,l)}^{n\ 0} = \gamma_{(k)}^{n} = 0$\ as established earlier. The relation can then be rewritten as a recursion relation for $\gamma_{(k,l)}^{m\ 1}$\ with $0<m<N-2$, which is essentially identical to~(\ref{eq:condition1_spec}) \begin{equation}\label{eq:pairwise_recursion} \left. \gamma_{(k,l)}^{m\ 1} C \right|_{\tiny{P}} = - \sum_{n=0}^{m-1} \left. \gamma_{(k,l)}^{n\ 1} \frac{1}{(m-n)!} \frac{\partial^{m-n} C}{\partial x_k^{m-n}} \right|_{\tiny{P}} \ . \end{equation} In complete analogy with the proof of the auxiliary result~(\ref{eq:aux2}) (the only difference being that the base step is now $m=1$), it follows that for $0<m<N-2$ \[ \gamma_{(k,l)}^{m\ 1} = \left. \frac{ \gamma_{(k,l)}^{0\ 1} C}{m!} \frac{\partial^m }{\partial x_k^m} \left( \frac{1}{C} \right) \right|_{\tiny{P}} \ , \] when condition~(\ref{eq:coeffs2}) is evaluated for $\vec{j}$\ with $j_n = m\delta_{kn} + \delta_{ln}$, this leads to \[ \left. \frac{ \gamma_{(k,l)}^{0\ 1} }{(N-2)!} \frac{\partial^{N-2} }{\partial x_k^{N-2}} \left( \frac{1}{C} \right) \right|_{\tiny{P}} = 0 \ . \] Once again, our assumption forces us to conclude that $\gamma_{(k,l)}^{0\ 1}=0$\ for arbitrary $l$.\\ \ \\ We have now established that $\gamma^{\vec{i}} =0$\ for all $|\vec{i}|=1$. The inductive extension of this result to $|\vec{i}|<N-1$\ is straightforward with $C|_{\tiny{P}} \neq 0$. Specifically, assume $\gamma^{\vec{i}} =0$\ for all $|\vec{i}| \leq n$. We can rewrite~(\ref{eq:coeffs1}) for any $\vec{j}$\ such that $|\vec{j}|=n+1$\ to obtain \[ \left. \gamma^{\vec{j}} C \right|_{\tiny{P}} = -\sum_{0< \vec{i} < \vec{j}} \left. \gamma^{\vec{i}} \frac{1}{(\vec{j} - \vec{i})!} \frac{\partial^{|\vec{j} - \vec{i}|} C}{\partial x_1^{j_1-i_1} \ldots \partial x_{M}^{j_M-i_M}} \right|_{\tiny{P}} \ . \] As the sum on the right is over $\vec{i}$\ of degree $n$\ and lower, every term vanishes by the inductive assumption and $\gamma^{\vec{i}}=0$\ follows for all $|\vec{i}|<N-1$. To eliminate $\gamma^{\vec{i}}$\ with $|\vec{i}|=N-1$\ we are forced to use condition~(\ref{eq:coeffs2}) instead and complete the inductive proof in the same manner as in the previous proof, i.e. by first proving $\gamma^{(N-1)\vec{v}_{(k)}}=0$\ then using steps to eliminate other coefficients of degree $N-1$. This completes the proof.\\ \ \\ A slight additional restriction on linear dependence immediately follows.\\ \ \\ \textbf{Corollary}: \emph{there is no open neighborhood in $\Gamma_Q$\ in which the set of (truncated) gradients $\{ d_{\varepsilon}\tilde{C}_{\vec{i}} \}_{1\leq |\vec{i}|\leq N-1}$\ is everywhere linearly dependent.}\\ \ \\ This follows since $\left. \frac{\partial C}{\partial x_k} \right|_{\tiny{P}} \neq 0$\ extends to some open neighborhood of $P$\ by continuity, while $\frac{\partial^{N-2} }{\partial x_k^{N-2}}\left(\frac{1}{C} \right) = 0$\ at best defines a hypersurface of codimension 1.\\ \ \\ At this stage, one may hope that the results may be strengthened further, however, unless additional restrictions are imposed, it is possible to manufacture simple low-order polynomial constraints that allow the gradients $\{ d_{\varepsilon}\tilde{C}_{\vec{i}} \}_{1\leq |\vec{i}|\leq N-1}$\ to be linearly dependent on some closed subsets of the values taken by $x_i$. Below we demonstrate this by an explicit example. Consider the algebra generated by a single element $\hat{x}$\ with the following constraint \[ \hat{C} = (\hat{x}-\hat{\mathbf{1}})(\hat{x}^2+\hat{\mathbf{1}}) \ . \] The classical constraint has a single real root $x=1$, which we will treat as the classical constraint surface. The constraint is regular for all real values of $x$\ as $\frac{d}{dx}C=2x^2+(x-1)^2$. Since the algebra is abelian the constraints have no reordering terms and $C_{\vec{i}} = \tilde{C}_{\vec{i}}$. Suppose we truncate the system at order $N=4$. Then, by the results established above, the gradients $\{ d_{\varepsilon}\tilde{C}_{\vec{i}} \}_{1\leq |\vec{i}|\leq 3}$\ are guaranteed to be linearly independent for any real value of $x$\ so long as $\frac{d^2}{dx^2} \left( \frac{1}{C} \right) \neq 0$. It is not difficult to see that \[ \frac{d^2}{dx^2} \left( \frac{1}{C} \right) = \frac{1}{C^3} \left( 4x^2(3-4x+3x^2) \right) \ , \] which vanishes at $x=0$. Indeed, one can verify directly that at this order \[ \left. \left( d_{\varepsilon}C_1 + d_{\varepsilon}C_2 \right) \right|_{x=0} =0 \ , \] where $C_i = \langle \widehat{\Delta x}^i \hat{C} \rangle$. At the point where, in addition to $x=0$, all the moments vanish as well we have $d_Q C_i|_{x=0,\ \varepsilon_j=0} = d_{\varepsilon}C_i$\ for $i>1$\ and we conclude \[ \left. \left( d_QC_1 + d_QC_2 \right) \right|_{x=0,\ \varepsilon_j=0} =0 \ , \] and the truncated set of constraints is truly functionally dependent at this point. In particular, in this example the failure of $\{ d_{\varepsilon}\tilde{C}_{\vec{i}} \}_{1\leq |\vec{i}|\leq 3}$\ to be everywhere linearly independent is not remedied when we take into account the gradients with respect to the expectation values, which vanish when moments are set to zero, nor can it be fixed by accounting for the reordering terms proportional to $\hbar$, which in this example do not appear. Instead, the situation may be helped in two ways: \begin{itemize} \item By noticing that $x=0$, where the constraint conditions become degenerate, is likely ``too far'' from the classical constraint surface at $x=1$, and so we would not expect the semiclassical truncation to be accurate for the expectation values lying so far from $C=0$\ surface, \item By noticing that for a polynomial $C$, and for some value $x_0$\ the condition \[ \left. \frac{d^{N-2}}{dx^{N-2}} \left( \frac{1}{C} \right) \right|_{x=x_0} = 0 \ , \] cannot hold for all $N$\ above a given order, thus truncating at successively higher orders one eventually arrives at the set of truncated constraint conditions that are functionally independent at $x_0$. In this example, for instance $\left.\frac{d^3}{dx^3} \left( \frac{1}{C} \right)\right|_{x=x_0}=0$\, but going one order higher $\left.\frac{d^4}{dx^4} \left( \frac{1}{C} \right)\right|_{x=x_0}=-24 \neq 0$, which guarantees independence at $x=0$\ for truncation at $N=6$. This, of course, has the caveat that there may now be other points at which the new set of conditions is not independent.\\ \ \\ \end{itemize} Finally, we would like to consider whether order $\hbar$\ contributions to $d_{\varepsilon} C_{\vec{i}}$\ can spoil linear independence. This can happen if \[ \sum_{1\leq|\vec{i}|\leq N-1} \left. \gamma^{\vec{i}} d_{\varepsilon} \tilde{C}_{\vec{i}} \right|_{\tiny{P}} = O(\hbar) \ , \] for coefficients $\gamma^{\vec{i}}$\ that are themselves of semiclassical order $0$. Taking these terms into account, on the classical constraint surface, we write relation~(\ref{eq:condition2_spec}) for $m=2$\ as \[ \left. \gamma_{(k)}^1 \frac{\partial C}{\partial x_k} \right|_{\tiny{P}} =O(\hbar) \ , \] which implies $\left. \frac{\partial C}{\partial x_k} \right|_{\tiny{P}} =O(\hbar)$. Similarly, off the constraint surface, the recursion relation~(\ref{eq:condition1_spec}) holds to order $\hbar$\ and hence \[ \gamma_{(k)}^m = \left. \frac{\gamma_{(k)}^1 C}{(m-1)!} \frac{\partial^{m-1}}{\partial x_k^{m-1}} \left( \frac{1}{C} \right) \right|_{\tiny{P}} + O(\hbar) \ , \] and we get the condition \[ \left. \frac{\gamma_{(k)}^1 C}{(N-2)!} \frac{\partial^{m}}{\partial x_k^{m}} \left( \frac{1}{C} \right) \right|_{\tiny{P}} = O(\hbar) \ . \] Thus, unless $\frac{ C}{(N-2)!} \frac{\partial^{m}}{\partial x_k^{m}} \left( \frac{1}{C} \right)$\ is `almost' zero, the gradients $\{d_{\varepsilon} C_{\vec{i}}\}_{1\leq|\vec{i}|\leq N-1}$\ inherit linear independence from their symmetric counterparts $\{ d_{\varepsilon}\tilde{C}_{\vec{i}} \}_{1\leq |\vec{i}|\leq N-1}$. \section{Conclusion} So what \emph{is} a semiclassical state on a Lie algebra according to our construction? In a nutshell, it is a linear functional on the corresponding universal enveloping algebra that assigns ``increasingly small'' (as powers of $\hbar$) values to increasingly high order generalized moments of the generating elements (defined by~(\ref{eq:moments})). This definition is consistent with the physical intuition that a semiclassical wavefunction is ``sharply peaked'' about some classical state. What use do we envisage for these semiclassical states? Assuming that a given quantum system is in a semiclassical state, truncating its space of states at some finite order in $\hbar$\ can dramatically simplify explicit computations of its physical properties. What remains is a truncated ``quantum phase space'', which, as demonstrated by the detailed analysis of Section~\ref{sec:PBOrder}, is equipped with a well-defined truncated ``quantum Poisson bracket'', representing the Lie bracket of the original Lie algebra. This bracket is of direct relevance to the dynamical evolution of quantum mechanics (see~(\ref{eq:Qevolution}) and the preceding discussion)---the truncated set of quantum dynamical equations is typically much simpler to solve numerically than the full Schr\"odinger equation (for an example see~\cite{EffCosm}). For the type of quantum systems considered here (see Section~\ref{sec:QLieAlg}) central elements of the universal enveloping algebra represent redundancy conditions that must be imposed in the form of constraints. Section~\ref{sec:Constraints} explicitly demonstrates the consistency of reducing the constraint system with the use of the semiclassical truncation in the case of a single generator of the center. A related question, not treated in the present work (but see~\cite{Casimir}), is the order-by-order characterization of the full set of inequalities that the positivity condition $\langle \hat{a} \hat{a}^* \rangle\geq 0$\ imposes on the generalized moments. In many situations of physical interest one needs to enforce these conditions, however, it remains to be demonstrated that, analogously to the constraints, the resulting set of inequalities is consistently reduced when the quantum system is truncated. Overall, we have shown that the effective equations based on the semiclassical approximation previously developed for the canonical quantum systems in~\cite{EffAc} and~\cite{EffEq2}, can be applied to systems based on a finite-dimensional Lie algebra by adding certain constraint conditions. This generalization, already employed in~\cite{Casimir}, has direct applications to quantum mechanics, quantum cosmology and gauge fields. \section*{Acknowledgements} The author would like to thank Martin Bojowald for encouragement, numerous helpful discussions, and comments on the draft version of this manuscript.
\section{Introduction} V4334 Sgr (a.k.a.\ Sakurai's object) is the central star of an old planetary nebula (PN) that underwent a very late thermal pulse (VLTP) a few years before its discovery in 1996 \citep{Na96}. During the VLTP it ingested its remaining hydrogen-rich envelope into the helium-burning shell and ejected the processed material shortly afterwards to form a new, hydrogen-deficient nebula expanding at a velocity of approximately 300~km\,s$^{-1}$ inside the old PN. The star brightened considerably and became a very cool, born-again asymptotic giant branch star with a spectrum resembling a carbon star. After a few years, dust formation started in the new ejecta and the central star became highly obscured. Emission lines were discovered: first He\,{\sc i} 10830~\AA\ in 1998 \citep{Ey99}, later in 2001 also optical forbidden lines from neutral and singly ionized nitrogen, oxygen, and sulfur, as well as very weak H$\alpha$ \citep{Ke02}. The distance to V4334 Sgr is poorly known, but is likely around 3 -- 4~kpc. \section{Evolutionary Models} Sakurai's object baffled the scientific community with its very fast evolution, must faster than pre-discovery models predicted. Three evolutionary models have been proposed to explain the fast evolution, all focusing on the hydrogen ingestion flash (HIF) in the helium burning shell. \citet{He01} and \citet{LM03} assume that hydrogen burning takes place close to the stellar surface due to the suppression of convection by the HIF. This is investigated further by Herwig using full 3D hydro models \citep{He11, He14}. These models show that the hydrogen ingestion proceeds through a global non-radial instability, which facilitates the transition from one spherically symmetric state into another one, and could never be computed in 1D models. \citet{LM03} were the first to predict the double-loop evolution in the HR diagram, later confirmed by Herwig's model in \citet{Ha05}. \citet{MB06} claim that they can reproduce very fast evolution by using very small time steps, but without changing the mixing physics. All of these models could be improved by constraining them with the temporal evolution of the stellar temperature. This was reasonably straightforward when the central star was still directly observable, but is much more difficult now that the star is heavily obscured. \articlefigure{flux_evol2.eps}{evol}{The evolution of the flux of various selected lines as a function of time.} \section{Optical Observations} We have been monitoring the evolution of the optical emission line spectrum since 2001 using low-resolution spectra taken with FORS1 and FORS2 on the ESO-VLT. The goal of this monitoring program is to derive the central star temperature as a function of time. First progress reports can be found in \citet{vH07, vH08}. The optical lines initially showed an exponential decline in intensity, and also a decreasing level of excitation. This trend continued until 2007. Between 2001 and 2007 the optical spectrum is consistent with a shock that occurred before 2001, and started cooling and recombining afterwards. The low electron temperature derived from the [N\,{\sc ii}] lines in 2001 (3200 -- 5500~K) and the [C\,{\sc i}] lines in 2003 (2300 -- 4300 K) is consistent with this \citep{vH07}. The earliest evidence for this shock is the detection of the He\,{\sc i} 10830~\AA\ recombination line in 1998 \citep{Ey99}. This line was absent in 1997. The shock must have occurred around 1998 and must have stopped soon after, leaving cooling and recombining gas in its wake. All line fluxes have been increasing since 2008! This is shown in Fig.~\ref{evol} for some selected strong lines. The spectrum we observed in 2013 is shown in Fig.~\ref{spec}. This confirms the trend for the [C\,{\sc i}] 9824 and 9850~\AA\ doublet reported by \citet{HJ14}. There are two exceptions: [O\,{\sc i}] 6300~\AA\ already started increasing in 2007 and the [N\,{\sc i}] 5198 and 5200~\AA\ doublet still decreased in flux in 2008. However, these exceptions may not be real as the lines in question suffer from strong telluric contamination. Also note that there is a strong discontinuous jump in the [O\,{\sc ii}] flux in 2008. \begin{figure}[!ht] \centering \leavevmode \includegraphics[width=.95\textwidth]{sakplot1.eps}{} \includegraphics[width=.95\textwidth]{sakplot2.eps}{} \caption{The FORS2 spectrum observed in 2013. The top panel shows the complete spectrum, and the bottom panel shows a zoom-in on the weaker lines in the same spectrum.\label{spec}} \end{figure} Here we report the first detection of several helium lines in 2008 (He\,{\sc i} 5876 and 7065~\AA) and 2009 (He\,{\sc i} 6678~\AA). Since 2010 the [S\,{\sc ii}] 6716 and 6731~\AA\ doublet is detected again. This is the first time this doublet is seen since 2001. This is due to a better signal-to-noise ratio in the spectra and the fact that the lines are brightening again. \citet{Ke02} detected H$\alpha$ at 5\% of the strength of [N\,{\sc ii}] 6583~\AA. Our spectra confirm this detection with a slightly greater strength of 7\% of the [N\,{\sc ii}] line. \section{Discussion} \citet{HJ14} believe that the reheating of the central star has started, mainly based on an increase in the blackbody temperature of the dust. This suggests that an onset of photoionization could be the cause of the rising line fluxes. However, in the most recent VLA observations V4334 Sgr was only barely detected at around 50~$\mu$Jy in 2012 and around 150~$\mu$Jy in 2013, indicating that the radio flux must have dropped since the detections presented in \citet{Ha05} and \citet{vH07, vH08}. The last value they reported was 550~$\mu$Jy in 2007. This seems inconsistent with an onset of photoionization. Alternatively, the sudden jump in the [O\,{\sc ii}] flux in 2008 could point to a second shock as the cause of the rising fluxes and we will adopt this as our working hypothesis. \citet{Ke02} observed two components in the shock-excited [N\,{\sc ii}] 6548 and 6583~\AA\ lines, one at a radial velocity of $-350$~km\,s$^{-1}$ and one at $+200$~km\,s$^{-1}$ relative to the central star. The blueshifted component was much stronger, resulting in the fact that the unresolved lines appeared blueshifted in the FORS spectra at about $-270$~km\,s$^{-1}$ relative to the central star. However, since 2011 we see a clear shift in radial velocity and the emission lines now appear at about $-120$~km\,s$^{-1}$ relative to the central star. This indicates that the emission lines seen up to 2007 came from a different region than those seen later on. Doing spectro-astrometry on the recent spectra indicates that the emission comes from a region less than 100~mas in size in the EW direction. \citet{HJ14} find a larger extent in the (roughly) NS direction. This suggests that the bipolar structure seen by \citet{Ch09} and \citet{HJ14} could be the origin of the shock emission. \articlefiguretwo{ratio1.eps}{ratio2.eps}{ratio}{Left panel -- two temperature sensitive line ratios as a function of time. Right panel -- two line ratios of ionized over neutral species as a function of time.} In Fig~\ref{ratio} we show various line ratios as a function of time. Note that these ratios are not corrected for extinction! The left panel shows two diagnostic ratios that are sensitive to electron temperature. Both indicate a decrease in electron temperature since 2008, assuming that the extinction is not increasing over time. This is inconsistent with photoionization and points to the presence of a shock. The line ratios suggest however that the electron temperature has been constant since 2012. The right-hand panel of Fig.~\ref{ratio} shows line ratios of ionized over neutral species from nitrogen and oxygen. These ratios are more difficult to interpret. A change in density would affect the degree of ionization in the gas, and if the density is high enough, the line ratios would also be affected by a changing degree of collisional de-excitation. The latter doesn't appear to be a problem as the observed [S\,{\sc ii}] line ratio indicates an electron density well below the critical density of all lines. The gas density could be dropping due to expansion of the nebula, or could be rising due to compression by the shock. Both line ratios indicate a drop in the degree of ionization (assuming that the extinction is not decreasing) at least since 2012. This is inconsistent with photoionization and would point to either the strength of the shock diminishing or the density of the gas increasing. Either interpretation would point to the presence of a shock. If the central star temperature is indeed rising, as suggested by \citet{HJ14}, it could have caused an increase in the mass loss and wind velocity from the central star, which is now causing shock emission in the bipolar structure. The sudden jump in the [O\,{\sc ii}] flux, as well as the dropping electron temperature point to a shock. The rising flux could be due to the fact that the interaction area between wind and the bipolar structure is still growing and is now approaching a maximum. We would like to emphasize that the analysis of our data is not yet complete, and all scenarios presented here are preliminary. \acknowledgements PvH acknowledges support from the Belgian Science Policy Office through the ESA PRODEX program.
\section{Introduction} \label{sec:intro} In an incomplete financial setting with noise governed by a continuous martingale and in which the investor's preferences are modeled by a negative power utility function, we provide a second-order Taylor expansion of the investor's value function with respect to perturbations of the underlying market price of risk process. We show that tractable models can be used to approximate highly intractable ones as long as the latter can be interpreted as perturbations of the former. As a by-product of our analysis we explicitly construct first-order approximations of both the primal and the dual optimizers. Finally, we apply our approximation in two numerical examples. There are two different ways of looking at our contribution: as a tool to approximate the value function and perform numerical computations, or as a stability result with applications to statistical estimation. Let us elaborate on these, and the related work, in order. \par{\em An approximation interpretation.} The conditions for existence and uniqueness of the investor's utility optimizers are well-established (see \cite{KLSX} and \cite{KraSch99}). However, in general settings, the numerical computation of the investor's value function remains a challenging problem. Various existing approaches include: \begin{enumerate} \item In Markovian settings, the value function can typically be characterized by a HJB-equation. Its numerical implementation through a finite-grid approximation is naturally subject to the curse of dimensionality. Many authors (see \cite{KimOmb96}, \cite{Wac2002}, \cite{ChaVic2005}, \cite{Kra05}, and \cite{Liu2007}) opt for affine and quadratic models for which closed-form solutions exist. Going beyond these specifications in high-dimensional settings by using PDE-techniques seems to be very hard computationally. \item In general (i.e., not necessarily Markovian) complete models, \cite{CviGouZap2003} and \cite{DetGarRin2003} provide efficient Monte Carlo simulation techniques based on the martingale method for complete markets developed in \cite{CoxHuang1} and \cite{KLS}. \item Other approximation methods are based on various Taylor-type expansions. The authors of \cite{Cam1993} and \cite{CamVic1999} log-linearize the investor's budget constraint as well as the investor's first-order condition for optimality. \cite{KogUpp2000} expand in the investor's risk-aversion coefficient around the log-investor (the myopic investor's problem is known to be tractable even in incomplete settings). When solving the HJB-equation numerically (using Longstaff-Schwartz type of techniques) \cite{BraGoySanStr2005} expand the value function in the wealth variable to a forth degree Taylor approximation. \item Based on the duality results in \cite{KLSX}, \cite{HauKogWan2006} provide an upper bound on the error stemming from using sub-optimal strategies. \cite{BicKraMun2013} propose a method based on minimizing over a subset of dual elements. This subset is chosen such that the corresponding dual utility can be computed explicitly and transformed into a feasible primal strategy. \item \revised{It is also important to mention the recent explosion in research in asymptotic methods in a variety of different ares in mathematical finance (transaction costs, pricing, etc.). Since we focus on model expansion in utility maximization in this paper, we simply point the reader to some of the most recent papers, namely \cite{AltMuhSon15}, and \cite{KalMuh15}, and the references therein, for further information.} \end{enumerate} In our work, no Markovian assumption is imposed and we deal with general, possibly incomplete, markets with continuous price processes. We note that while our results apply only to $p<0$, it is possible to extend them to $p\in(0,1)$ at the cost of imposing additional integrability requirements. We do not pursue such an extension; the parameter range $p\in (0,1)$ which we leave out seems to lie outside the typical range of risk-aversion parameters observed in practice (see, e.g., \cite{Szp86}). Moreover, we do not consider utility functions more general than the powers. While there are no significant additional mathematical difficulties in treating the general case under appropriate conditions on the relative risk-aversion coefficients, we do not believe that the added value justifies the corresponding notational and technical overhead. For example, all our results would become dependent on the agent's initial wealth, and this dependence would permeate the entire analysis. \par{\em A stability interpretation.} As we mentioned above, our contribution can also be seen as a stability result. It is well-known (see, e.g., \cite{Rogers}) that even in Samuelson's model, estimating the drift is far more challenging than estimating the volatility. \cite{LarZit07} identify the kinds of perturbations of the market price of risk process under which the value function behaves continuously. In the present paper we take the stability analysis one step further and provide a first-order Taylor expansion in an infinite-dimensional space of the market price of risk processes. This way, we not only identify the ``continuous'' directions, but also identify those features of the market price of risk process that affect the solution of the utility maximization problem the most (at least locally). Any statistical procedure which is performed with utility maximization in mind should, therefore, focus on those, salient, features in order to use the scarce data most efficiently. Similar perturbations have been considered by \cite{Monoyios}, but in a somewhat different setting. \cite{Monoyios} is based on Malliavin calculus and produces a first-order expansion for the utility-indifference price of an exponential investor in an It\^o-process driven market; some of the ideas used can be traced to the related work \cite{Dav06}. \par{\em Mathematical challenges.} From a mathematical point of view, our approach is founded on two ideas. One of them is to extend the techniques and results of \cite{LarZit07}; indeed, the basic fact that the optimal dual minimizers converge when the market prices of risk process does is heavily exploited. It does not, however, suffice to get the full picture. For that, one needs to work on the primal and the dual problems simultaneously and use a pair of bounds. The ideas used there are related to and can be interpreted as a nonlinear version of the primal-dual second-order error estimation techniques first used in \cite{Hen2002} in the context of mathematical finance. The first-order expansion in the quantity of the unspanned contingent claim developed in \cite{Hen2002} was generalized in \cite{KramkovSirbu2006b} (see also \cite{KramkovSirbu2006a}). The arguments in these papers rely on convexity and concavity properties in the expansion parameter (wealth and number of unspanned claims). This is not the case in the present paper; indeed, when seen as a function of the underlying market price of risk process, the investor's value function is neither convex nor concave and a more delicate, local, analysis needs to be performed. \bigskip \par{\em Numerical examples.} In Section \ref{sec:examples} we use two examples to illustrate how our approximation performs under realistic conditions. First, we consider the Kim-Omberg model (see \cite{KimOmb96}) which is widely used in the financial literature. Under a calibrated set of parameters, we find that our approximation is indeed very accurate when compared to the exact values. Our second example belongs to a class of extended affine models introduced in \cite{CheFilKim07}. The authors show that this class of models has superior empirical properties when compared to popular affine and quadratic specifications (such as those used, e.g., in \cite{Liu2007}). The resulting optimal investment problem for the extended affine models, unfortunately, does not seem to be explicitly solvable. Our approximation technique turns out to be easily applicable and our error bounds are quite tight in the relevant parameter ranges. \section{A family of utility-maximization problems} \label{sec:problem} \subsection{The setup} \label{sse:setup} We work on a filtered probability space $(\Omega,\sF, {\mathbb F}= \prf{\sF_t}, {\mathbb P})$, with the finite time horizon $T>0$. We assume that the filtration ${\mathbb F}$ is right-continuous and that the $\sigma$-algebra $\sF_0$ consists of all ${\mathbb P}$-trivial subsets of $\sF$. Let $M$ be a continuous local martingale, and let $\Re$, $\varepsilon\geq 0$ be a family of continuous ${\mathbb F}$-semimartingales given by \begin{equation} \label{equ:7283} \begin{split} \Re := M + \int_0^{\cdot} \upeps{\ld}_t\, d\ab{M}_t, \text{ on } [0,T],\text{ where } \upeps{\ld}: = \lambda + \varepsilon \lambda', \end{split} \end{equation} for a pair $\lambda,\lambda'\in\sP_{M}^2$, where $\sP_{M}^2$ denotes the collection of all progressively measurable processes $\pi$ with $\int_0^T \pi_t^2\, d\ab{M}_t<\infty$. As $\upeps{S} := {\mathcal E}(\Re)$ (where ${\mathcal E}$ denotes the stochastic exponential) will be interpreted as the price process of a financial asset, the assumption that $\upeps{\ld}\in\sP_{M}^2$ can be taken as a minimal no-arbitrage-type condition. We remark right away that further integrability conditions on $\lambda$ and $\lambda'$ will need to be imposed below for our main results to hold. \subsection{The utility-maximization problem} Given $x>0$ and $\varepsilon \in [0,\infty)$, let $\upeps{\sX}(x)$ denote the set of all nonnegative wealth processes starting from initial wealth $x$ in the financial market consisting of $\upeps{S} := {\mathcal E}(\Re)$ and a zero-interest bond, i.e., \[ \upeps{\sX}(x) := \Bsets{x {\mathcal E}\big(\textstyle \int_0^T \pi_t\, d\Re_t\big)}{ \pi \in\ \sP_{M}^2}.\] Here, $\pi$ is interpreted as the fraction of wealth invested in the risky asset $\upeps{S}$. The investor's preferences are modeled by a CRRA (power) utility function with the risk-aversion parameter $p<0$: \begin{equation}\label{equ:U} U(x) := \frac{x^p}{p}, \quad x>0. \end{equation} The value function of the corresponding optimal-investment problem is defined by \begin{equation}\label{equ:ue} \upeps{u}(x) := \sup_{X\in\upeps{\sX}(x)} {\mathbb E}[ U(X_T)],\ x>0. \end{equation} \subsection{The dual utility-maximization problem} As is usual in the utility-ma\-xi\-mi\-za\-ti\-on literature, a fuller picture is obtained if one also considers the appropriate version of the optimization problem dual to \eqref{equ:ue}. For that, we need to examine the no-arbitrage properties of the set of models introduced in Section \ref{sse:setup} above. We observe, first, that the assumptions we placed on the market price of risk processes $\upeps{\ld}$ above are not sufficient to guarantee the existence of an equivalent martingale measure (NFLVR). They do preclude so-called ``arbitrages of the first kind'' and imply the related condition NUBPR. In particular, for all $x,y>0$ and $\varepsilon\geq 0$ there exists a (strictly) positive c\` adl\` ag supermartingale $Y$ with the property that $Y_0=y$ and $YX$ is a supermartingale for each $X\in\upeps{\sX}(x)$; we denote the set of all such processes by $\upeps{\sY}(y)$. While this is a consequence of the condition NUBPR in general, in this case an example of a process in $\upeps{\sY}(y)$ is given, explicitly, as $y\upeps{Z}$, where $\upeps{Z}$ is the \emph{minimal} local martingale density: \begin{equation} \label{equ:Ze} \begin{split} \upeps{Z} = {\mathcal E}(-\int_0^{\cdot} \upeps{\ld}_t\, dM_t). \end{split} \end{equation} Having described the dual domain, we remind the reader that the \emph{conjugate} utility function $V:(0,\infty)\to\R$ is defined by \begin{equation}\label{equ:V} V(y) := \sup_{x>0} \left( U(x) - xy\right) = \frac{y^{-q}}{q},\text{ where } q:=\tfrac{p}{1-p}\in(-1,0). \end{equation} We define the \emph{dual value function} $\upeps{v}:(0,\infty)\to \R$ by \begin{equation} \label{equ:ve} \begin{split} \upeps{v}(y) := \inf_{Y\in \upeps{\sY}(y) } {\mathbb E}[ V(Y_T)],\ y>0,\ \varepsilon\geq 0. \end{split} \end{equation} Due to negativity (and, a fortiori, finiteness) of the primal value function $\upeps{u}$, the (abstract) Theorem 3.1 of \cite{KraSch99} can now be applied (see also \cite{Mostovyi2011}). Its main assumption, namely the bipolar relationship between the primal and dual domains, holds due to the existence of the num\' eraire process, given explicitly by $1/\upeps{Z}$ (see Theorem 4.12~in \cite{KarKar07}). One can also use a simpler argument (see \cite{Lar11}), which applies only to the case of a CRRA utility with $p<0$, to obtain the following conclusions for all $\varepsilon\geq 0$: \begin{enumerate} \item both $\upeps{u}$ and $\upeps{v}$ are finite and the following conjugacy relationships hold \begin{equation} \label{equ:conj} \begin{split} \upeps{v}(y) = \sup_{x>0} \Big( \upeps{u}(x)-xy \Big),\text{ and } \upeps{u}(x) = \inf_{y>0} \Big( \upeps{v}(y)+xy \Big). \end{split} \end{equation} \item For all $x,y>0$ there exist optimal solutions $\upeps{\hat{X}}(x)\in\upeps{\sX}(x)$ and $\upeps{\hat{Y}}(y)\in\upeps{\sY}(y)$ of \eqref{equ:ue} and \eqref{equ:ve}, respectively, and are related by \[ U'(\upeps{\hat{X}}_T(x)) = \upeps{\hat{Y}}_T (\upeps{y}(x)) \text{ where } \upeps{y}(x) = \trn{}{x} \upeps{u}(x) = p x^{p-1} \upeps{u}(1). \] \item \revised{The product $\upeps{\hat{X}}\upeps{\hat{Y}}$ is a uniformly-integrable martingale. In particular \[ {\mathbb E}[ \upeps{\hat{X}}_T \upeps{\hat{Y}}_T]=xy.\]} \end{enumerate} The homogeneity of the utility function $U$ and its conjugate $V$ transfers to the value functions $\upeps{u}$ and $\upeps{v}$ and the optimal solutions $\upeps{\hat{X}}$ and $\upeps{\hat{Y}}$: \begin{equation} \label{equ:simpl} \begin{split} \upeps{u}(x) &= x^p \upeps{u},\quad \upeps{v}(y)=y^{-q} \upeps{v},\\ \upeps{\hat{X}}(x)&=x \upeps{\hat{X}},\quad \upeps{\hat{Y}}(y)=y \upeps{\hat{Y}}, \end{split} \end{equation} where, to simplify the notation, we write $\upeps{u}, \upeps{v}, \upeps{\hat{X}}$ and $\upeps{\hat{Y}}$ for $\upeps{u}(1)$, $\upeps{v}(1)$, $\upeps{\hat{X}}(1)$ and $\upeps{\hat{Y}}(1)$, respectively. \subsection{A change of measure} For $\varepsilon=0$ we denote by $\upz{\hat{\pi}}$ the primal optimizer, i.e., the process in $\sP_{M}^2$ such that \[ \upz{\hat{X}} = {\mathcal E}( \int_0^{\cdot} \upz{\hat{\pi}}_u\, d\upz{R}_u).\] We define the probability measure $\upz{\tilde{\PP}}$ by \begin{equation} \label{equ:Girsanov} \begin{split} \rn{\upz{\tilde{\PP}}}{{\mathbb P}} = \upz{\hat{X}}_T \upz{\hat{Y}}_T\, \Big(= \oo{\upz{v}} V(\upz{\hat{Y}}_T) = \oo{\upz{u}} U(\upz{\hat{X}}_T)\Big), \end{split} \end{equation} \revised{where the last two equalities follow from the identities $x U'(x) = p U(x)$ and $ y V'(y) = -q V(y)$, and the relations between the value functions outlined above.} The measure $\upz{\tilde{\PP}}$ has been in the mathematical finance literature for a while (see, e.g., p. 911-2 in \cite{KraSch99}). The explicit form of $\upz{\tilde{\PP}}$ is not generally available, but, we note that, by Girsanov's Theorem (see \eqref{equ:ZENL} and the discussion around it), the process \begin{equation} \label{equ:tMp} \begin{split} \tM^p := M + \int_0^{\cdot} \Big( \lambda_t - \upz{\hat{\pi}}_t \Big)\, d\ab{M}_t \end{split} \end{equation} is a $\upz{\tilde{\PP}}$-local martingale; \revised{this fact will be used below in the proof of Proposition \ref{pro:u-down}.} \section{The problem and the main results} We first provide first-order expansions and error estimates of the primal and dual value functions. Secondly, we provide an expansion of the optimal controls in the Brownian setting. \subsection{Value functions} At the basic level, we are interested in the first-order properties of the convergence, as $\varepsilon\searrow 0$, of the value functions of the problems $\upeps{u}$ and $\upeps{v}$ to the value functions $\upz{u}$ and $\upz{v}$ of the ``base'' model (corresponding to $\varepsilon=0$). To familiarize ourselves with the flavor of the results we can expect in the general case, we start by analyzing a similar problem for the logarithmic utility. It has the advantage that it admits a simple explicit solution. Let $\ue_{\log}(x)$ and $\ve_{\log}(y)$ denote the value function of the utility maximization problem as in \eqref{equ:ue} and \eqref{equ:ve} above, but with $U(x)=\log(x)$ and $V(y) = \sup_x (U(x)-xy)=-\log(y) -1$. It is a classical result that, as long as ${\mathbb E}[ \int_0^T (\lambda_t^2 + (\lambda'_t)^2)\, d\ab{M}_t]<\infty$, we have \[ \ue_{\log}(x) = \log(x)+\tot {\mathbb E}[ \int_0^T (\upeps{\ld}_t)^2\, d\ab{M}_t] \text{ and } \ve_{\log} = \ue_{\log} -1.\] The (exact) second-order expansion in $\varepsilon$ of $\ue_{\log}(x)$ is thus given by \begin{equation*} \label{equ:4AA5} \begin{split} \ue_{\log}(x) &= \uz_{\log}(x) + \varepsilon {\mathbb E}[ \int_0^T \lambda_t \lambda'_t\, d\ab{M}_t] + \tot \varepsilon^2 {\mathbb E}[ \int_0^T (\lambda'_t)^2\, d\ab{M}_t]\\ &=\uz_{\log}(x) + \varepsilon {\mathbb E}[ \int_0^T \lambda'_t\, d\upz{R}_t] + \tot \varepsilon^2 {\mathbb E}[ \int_0^T (\lambda'_t)^2\, d\ab{M}_t],\\ \end{split} \end{equation*} where $\upz{R}$ is defined in (\ref{equ:7283}). We cannot expect the value function to be a second order polynomial in $\varepsilon$ in the case of a general power utility. We do obtain a formally similar first-order expansion in Theorem \ref{thm:main1} below and an analogous error estimate in Theorem \ref{thm:main2}. Section 5 is devoted to their proofs. We remind the reader of the homogeneity relationships in \eqref{equ:simpl}; they allow us to assume from now on that $x=y=1$. \begin{theorem}[The G\^ateaux derivative] \label{thm:main1} In the setting of Section \ref{sec:problem}, we assume that \begin{equation} \label{equ:A1} \begin{split} \int_0^T(\lambda'_t)^2\, d\ab{M}_t \in \el^{1-p}({\mathbb P}) \text{ and } \int_0^T \lambda'_t\, d\upz{R}_t \in \cup_{s>(1-p)} \el^s({\mathbb P}). \end{split} \end{equation} Then, with $\upz{\Delta} := {\mathbb E}^{\upz{\tilde{\PP}}}[\textstyle \int_0^T \lambda'_t\, d\upz{R}_t]$, where $\upz{\tilde{\PP}}$ is defined by \eqref{equ:Girsanov}, we have \begin{align} \label{equ:main1-u} \derep{\upeps{u}}:=\lim_{\varepsilon \searrow 0} \oo{\varepsilon} \Big( \upeps{u} - \upz{u} \Big) &= p \upz{u} \upz{\Delta}, \text{ and } \\ \label{equ:main1-v} \derep{\upeps{v}}:=\lim_{\varepsilon \searrow 0} \oo{\varepsilon} \Big( \upeps{v} - \upz{v} \Big) &= q\upz{v} \upz{\Delta}.\qedhere \end{align} \end{theorem} \begin{theorem}[An error estimate] \label{thm:main2} In the setting of Section \ref{sec:problem}, we assume that \begin{equation} \label{equ:A2} \begin{split} \int_0^T(\lambda'_t)^2\, d\ab{M}_t, \int_0^T \lambda'_t\, d\upz{R}_t \in \el^{2(1-p)}({\mathbb P}) \text{ and } \Phi^2 e^{\varepsilon_0 |p| \Phi^-}\in\el^1(\upz{\tilde{\PP}}), \end{split} \end{equation} for some $\varepsilon_0>0$, where $\Phi := \int_0^T \upz{\hat{\pi}}_t \lambda'_t\, d\ab{M}_t$. Then there exist constants $C>0$ and $\varepsilon_0'\in (0,\varepsilon_0]$ such that for all $\varepsilon\in [0,\varepsilon'_0]$ we have \begin{align} \label{equ:main2-u} \Big| \upeps{u} - \upz{u} - \varepsilon p \upz{u} \upz{\Delta} \Big| & \leq C \varepsilon^2, \text{ and } \\ \label{equ:main2-v} \Big| \upeps{v} - \upz{v} - \varepsilon q \upz{v} \upz{\Delta} \Big| & \leq C \varepsilon^2. \end{align} \end{theorem} \begin{remark}\ \label{rem:761C} \begin{enumerate} \item It is perhaps more informative to think of the results in Theorems \ref{thm:main1} and \ref{thm:main2} on the logarithmic scale. As is evident from \eqref{equ:main1-u} and \eqref{equ:main1-v}, the functions $\upeps{u}$ and $\upeps{v}$ admit the right \emph{logarithmic} derivative $p\upz{\Delta}$ and $q\upz{\Delta}$, respectively, at $\varepsilon=0$. Moreover, we have the following small-$\varepsilon$ asymptotics: \[ \upeps{u} = \upz{u} e^{\varepsilon p \upz{\Delta} + O(\varepsilon^2)} \text{ and } \upeps{v} = \upz{v} e^{\varepsilon q \upz{\Delta} + O(\varepsilon^2)}.\] If one takes one step further and uses the \emph{certainty equivalent} $\upeps{\mathrm{CE}}$, given by \[ U(\upeps{\mathrm{CE}}) = \upeps{u},\] we note that $\upz{\Delta}$ is precisely the infinitesimal growth-rate of $\upeps{\mathrm{CE}}$ at $\varepsilon=0$ - an $\varepsilon$-change of the market price of risk in the direction $\lambda'$ yields to an $e^{\varepsilon \upz{\Delta}}$-fold increase in the certainty-equivalent of the initial wealth. \item \label{rem:main2} A careful analysis of the proof of Theorem \ref{thm:main2} below reveals the following, additional, information: \begin{enumerate} \item The proof of Proposition \ref{pro:u-down} reveals that $\upz{\Delta}={\mathbb E}^{\upz{\tilde{\PP}}}[\Phi]$. \item The condition involving $\Phi$ in \eqref{equ:A2} is needed only for the upper bound in \eqref{equ:main2-u} and the lower bound in \eqref{equ:main2-v}. The other two bounds hold for all $\varepsilon\geq 0$ even if \eqref{equ:A2} holds with $\varepsilon_0=0$. \item The constants $C$ and $\varepsilon_0'$ depend - in a simple way - on $\varepsilon_0$, $p$ and the $\el^{2(1-p)}(\upz{\tilde{\PP}})$- and $\el^1(\upz{\tilde{\PP}})$-bounds of the random variables in \eqref{equ:A2}. For two one-sided bounds, explicit formulas are given in Propositions \ref{pro:v-up} and \ref{pro:u-down}. The other two bounds are somewhat less informative so we do not compute them explicitly. The reader will find an example of how this can be done in a specific setting in Subsection \ref{sse:extaff}. \item Even though we cannot claim that the functions $\upeps{u}$ and $\upeps{v}$ are convex or concave, it is possible to show their local {\it semiconcavity} in $\varepsilon$ (see \cite{Can04}). This can be done via the techniques from the proof of Theorem~\ref{thm:main2}. \end{enumerate} \item The assumption of constant risk aversion (power utility) allows us to incorporate many stochastic interest-rate models into our setting. Indeed, provided that $c:= \ee{e^{p\int_0^T r_tdt}} <\infty$, we can introduce the probability measure ${\mathbb P}^r$, defined by \begin{align} \rn{{\mathbb P}^r}{{\mathbb P}}:= c e^{p\int_0^T r_t\, dt}, \end{align} on $\sF_T$. For any admissible wealth process $X$ we then have \[ {\mathbb E}[ U(X_T) ] = c\, {\mathbb E}^{{\mathbb P}^r}\left[ U\Big( X_T e^{-\int_0^T r_u\, du}\Big) \right].\] This way, the utility maximization under ${\mathbb P}^r$ with a zero interest rate becomes equivalent to the utility maximization problem under ${\mathbb P}$ with the interest rate process $\prf{r_t}$. \cite{Zitkovic2005} and \cite{Mostovyi2011} consider the setting of utility maximization with stochastic utility which embeds stochastic interest rates. Practical implementation of the above idea depends on how explicit one can be about the Girsanov transformation associated with ${\mathbb P}^r$. It turns out, fortunately, that many of the widely-used interest-rate models, such as Vasi\v cek, CIR, or the quadratic normal models (see, e.g., \cite{Mun13} for a textbook discussion of these models) allow for a fully explicit description (often due to their affine structure). For example, in the Vasi\v cek model, the Girsanov drift under ${\mathbb P}^r$ can be computed quite explicitly, due to the underlying affine structure. Indeed, suppose that $r$ has the Ornstein-Uhlenbeck dynamics of the form \[ dr_t := \kappa (\theta-r_t)\, dt + \beta\, dB_t,\ r_0\in\R,\] where $B$ is a Brownian motion and $\kappa>0$, $\theta,\beta\in\R$. Then the process \[ B^{(p)} := B - \int_0^{\cdot} b(T-t)\, dt,\text{ where } b(t) = \tfrac{\beta p}{\kappa} (1-e^{-\kappa t}),\] is a ${\mathbb P}^r$-Brownian motion. \end{enumerate} \end{remark} \subsection{Optimal controls} The estimates \eqref{equ:main2-u} and \eqref{equ:main2-v} are of type $O(\varepsilon^2)$. A slight adjustment to the below proof of Proposition \ref{pro:u-down} shows that the wealth process $\tilde{X} :=\mathcal{E}\big(\int\hat{\pi}^{(0)}dR^{(\varepsilon)})$ satisfies (see \ref{lower_bound_est}) \begin{align*} \big|{\mathbb E}[U(\tilde{X}_T)] - u^{(0)}(1+\varepsilon p \Delta^{(0)}) \big|&\le \frac12p^2\varepsilon^2 |u^{(0)}| {\mathbb E}^{\upz{\tilde{\PP}}}[\Phi^2e^{\varepsilon |p| \Phi^-}]. \end{align*} Therefore, under the conditions of Theorem \ref{thm:main2}, $\hat{\pi}^{(0)}$ is an $O(\varepsilon^2)$-optimal control for the $\varepsilon$-model because the triangle inequality produces a constant $C>0$ such that $$ \big|{\mathbb E}[U(\tilde{X}_T)] -u^{(\varepsilon)} \big| \le C \varepsilon^2, $$ for all $\varepsilon>0$ small enough. In this section we will provide a correction term to $\hat{\pi}^{(0)}$ such that the resulting wealth process upgrades the convergence to $o(\varepsilon^2)$. For simplicity, we consider the (augmented) filtration generated by $(B,W)$ where $B\in \R$ and $W\in\R^d$, $d\in\mathbb{N}$, are two independent Brownian motions. In \eqref{equ:7283} we take \begin{align}\label{dM_BM} dM_t := \sigma_t dB_t,\quad M_0:=0, \end{align} for a process $\sigma\in \sP_{B}^2$ with $\sigma\neq0$. We define $\upz{\tilde{\PP}}$ by \eqref{equ:Girsanov} and we denote by $(B^{\upz{\tilde{\PP}}},W^{\upz{\tilde{\PP}}})$ the corresponding $\upz{\tilde{\PP}}$-Brownian motions. Provided that $\Phi :=\int_0^T \upz{\hat{\pi}}_t \lambda'_t\sigma^2_t dt \in \el^{2}(\upz{\tilde{\PP}})$, $\Phi$ has the unique martingale representation under $\upz{\tilde{\PP}}$ \begin{align}\label{Itorep} \Phi = {\mathbb E}^{\upz{\tilde{\PP}}}[\Phi] + \int_0^T \gamma^B_{t}\sigma_tdB^{\upz{\tilde{\PP}}}_t + \int_0^T \gamma^W_{t} dW^{\upz{\tilde{\PP}}}_t, \end{align} where we have used $\sigma\neq0$. Because $\Phi\in \el^{2}(\upz{\tilde{\PP}})$ the two processes $\gamma^W$ and $\gamma^B$ in \eqref{Itorep} satisfy the integrability conditions $$ {\mathbb E}^{\upz{\tilde{\PP}}}\Big[\int_0^T\Big( (\gamma^B_{t}\sigma_t)^2 + (\gamma^W_t)^2\Big)dt\Big]<\infty. $$ These square integrability properties will be used in the proof of the next theorem. \begin{theorem}[2nd order expansion] \label{thm:main3} In the above Brownian setting, we assume \begin{align}\label{2nd_int} \int_0^T(\lambda'_t)^2\sigma^2_tdt \in \el^{1-p}({\mathbb P})\cap \el^{1}(\upz{\tilde{\PP}}) \text{ and } \int_0^T \hat{\pi}_t^{(0)} \lambda_t' \sigma_t^2 dt \in \el^{2}(\upz{\tilde{\PP}}), \end{align} as well as the existence of a constant $\varepsilon_0 >0$ such that $\delta :=\tfrac{\lambda'+p\gamma^B}{1-p}$ satisfies \begin{align}\label{exp_reg} e^{p\int_0^T\big (\varepsilon\hat{\pi}^{(0)} \lambda'+ \varepsilon^2 (\delta \lambda'-\frac12\delta^2 ) \big)\sigma^2dt + p\varepsilon \int_0^T\delta\sigma dB_t^{\upz{\tilde{\PP}}} }\in \el^{1}(\upz{\tilde{\PP}}), \end{align} for all $\varepsilon \in (0,\varepsilon_0)$. Then we have \begin{align} \label{equ:main3-u} \upeps{u} - \upz{u} - \varepsilon p \upz{u} \upz{\Delta} - \tfrac12 \varepsilon^2p\upz{u}\left(\Delta^{(00)}+p(\Delta^{(0)})^2 \right)& \in O(\varepsilon^3), \\ \label{equ:main3-v} \upeps{v} - \upz{v} - \varepsilon q \upz{v} \upz{\Delta} - \tfrac12 \varepsilon^2q\upz{v}\left(\Delta^{(00)}+q(\Delta^{(0)})^2 \right) & \in O(\varepsilon^3), \end{align} as $\varepsilon\searrow 0$. In \eqref{equ:main3-u} and \eqref{equ:main3-v} we have defined \begin{align} \Delta^{(00)} &:= {\mathbb E}^{\upz{\tilde{\PP}}}\left[ \int_0^T\left(p|\gamma^W_{t}|^2 +\frac{(\lambda'_t)^2 + p\gamma^B_t(\gamma^B_t+2\lambda'_t)}{1-p}\sigma^2_t\right)dt\right],\label{Delta00} \end{align} where the processes $\gamma^B$ and $\gamma^W$ are given by the martingale representation \eqref{Itorep}. \end{theorem} \begin{remark}\ \label{rem:controls} \begin{enumerate} \item The below proof of Theorem \ref{thm:main3} shows that the process \begin{align}\label{optimal_control_correction} \tilde{\pi} := \hat{\pi}^{(0)} + \varepsilon \tfrac{\lambda'+p\gamma^B}{1-p}, \end{align} is an $O(\varepsilon^3)$-optimal control for the $\varepsilon$-model in the sense that the wealth process $\tilde{X} :=\mathcal{E}\big(\int\tilde{\pi}dR^{(\varepsilon)})$ satisfies $$ {\mathbb E}[U(\tilde{X}_T)] -u^{(\varepsilon)} \in O(\varepsilon^3)\quad \text{as}\quad \varepsilon \searrow 0. $$ \item Because the filtration is generated by $(B,W)$, the optimizer $\hat{H}^{(0)}$ for the dual problem \eqref{equ:ve} can be written as $\hat{H}^{(0)} = \mathcal{E}(-\int \hat{\nu}^{(0)}dW)$ for a $d$-dimensional process $\hat{\nu}^{(0)}$ in $\sP_{W}^2$. The below proof of Theorem \ref{thm:main3} also shows that the process \begin{align}\label{optimal_dual_control_correction} \tilde{\nu} := \hat{\nu}^{(0)} - \varepsilon p\gamma^W, \end{align} is an $O(\varepsilon^3)$-optimal dual control in the $\varepsilon$-model. \item Throughout the paper we have considered $\varepsilon=0$ as the base model. Because we can write $$ \lambda + (\bar{\varepsilon} +\varepsilon)\lambda' = \lambda + \bar{\varepsilon}\lambda' +\varepsilon\lambda', $$ for any $\bar{\varepsilon}\in[\varepsilon_L,\varepsilon_U]$ with $\varepsilon_L<\varepsilon_U$, we can use Theorem \ref{thm:main3} for the base model $\lambda+ \bar{\varepsilon}\lambda'$ to provide a 2nd order Taylor expansion around any point $\bar{\varepsilon}$. Therefore, whenever $\Delta^{(0)}$ and $\Delta^{(00)}$ are bounded uniformly in $\bar{\varepsilon}\in[\varepsilon_L,\varepsilon_U]$, Theorem 3 in \cite{Oli54} ensures that $u^{(\varepsilon)}$ is twice differentiable in $\varepsilon$. \item An easy way of eliminating the stochastic $B^{\upz{\tilde{\PP}}}$-integral appearing in \eqref{exp_reg} is to use H\"older's inequality with the exponents $-1/q$ and $(1-p)$; see Section \ref{sse:extaff} below for an example. \end{enumerate} \end{remark} \section{Proofs of the main theorems} \label{sec:proofs} We start the proof with a short discussion of the special structure the dual domain $\upeps{\sY}$ has when the stock-price process $\upeps{S} = \mathcal{E}(\Re)$ is continuous. Indeed, it has been shown in \cite{LarZit07}, Proposition 3.2, p.~1653, that in that case the maximal elements in $\upeps{\sY}$ (in the pointwise order) are precisely local martingales of the form \begin{equation*} \begin{split} Y = \upeps{Z} H ,\ H\in\sH, \end{split} \end{equation*} where $\sH$ denotes the set of all $M$-orthogonal positive local martingales $H$ with $H_0=1$. We remark that even though the results in \cite{LarZit07} were written under the assumption NFLVR, a simple localization argument shows that they apply under the present conditions, as well. Hence, we can write \[ \upeps{v} = \inf_{H\in\sH} {\mathbb E}[ V(\upeps{Z}_T H_T)],\] and the minimizer $\upeps{\hat{Y}}$ always has the form \begin{equation} \label{equ:ZENL} \begin{split} \upeps{\hat{Y}} = \upeps{Z} \upeps{\hat{H}},\text{ for some $\upeps{\hat{H}}\in\sH$.} \end{split} \end{equation} Finally, we introduce two shortcuts for expressions that appear frequently in the proof: \begin{equation} \label{equ:etaLambda} \begin{split} \eta := \int_0^T \lambda'_t\, d\upz{R}_t,\ \Lambda := \int_0^T (\lambda'_t)^2\, d\ab{M}_t, \end{split} \end{equation} and remind the reader that $\Phi := \int_0^T \upz{\hat{\pi}}_t \lambda'_t\, d\ab{M}_t$ and $\Delta^{(0)} := {\mathbb E}^{\upz{\tilde{\PP}}}[ \eta]$. It will be useful to keep in mind that $(1-p)(1+q)=1$ and that $-1/q$ and $1-p$ are conjugate exponents. \subsection{A proof of Theorem \ref{thm:main1}} The proof is based on the stability results of \cite{LarZit07} and the following lemma: \begin{lemma} \label{lem:K-eps} Let $\set{\upeps{K}}_{\varepsilon\geq 0}$ be a family of positive random variables such that \begin{enumerate} \item ${\mathbb E}[ \updel{Z}_T \upeps{K}]\leq 1$ for all $\varepsilon,\delta \geq 0$, and \item $\upeps{K} \to \upz{K}$ in probability, as $\varepsilon \searrow 0$. \end{enumerate} Then, under the conditions of Theorem \ref{thm:main1}, we have \[ \lim_{\varepsilon\searrow 0} \oo{\varepsilon} {\mathbb E}\Big[ V(\upeps{Z}_T \upeps{K}) - V(\upz{Z}_T \upeps{K})\Big] = q \ee{ V(\upz{Z}_T \upz{K})\eta}.\] \end{lemma} \begin{proof} The map $\varepsilon\mapsto \upeps{Z}_T$ is almost surely continuously differentiable; \revised{ indeed, we have \[ \log(\upeps{Z}_T) = \log(\upz{Z}_T) - \varepsilon \int_0^T \lambda'_t\, d\upz{R}_t - \tot \varepsilon^2 \int_0^T (\lambda'_t)^2\, d\ab{M}_t, \] and, so, \[ \tfrac{d}{d \varepsilon} \upeps{Z}_T = -\upeps{Z}_T \Big( \eta + \varepsilon \Lambda \Big), \text{ a.s.} \] Therefore, } \begin{align} \label{equ:primal1} V(\upeps{Z}_T K) -V(\upz{Z}_T K) = \int_0^\varepsilon q V(\updel{Z}_T K) (\eta +\delta \Lambda)d\delta, \end{align} for each $\varepsilon$ and each positive random variable $K$. Thus, \begin{equation} \label{equ:378} \begin{split} V(\upeps{Z}_T \upeps{K}) & -V(\upz{Z}_T \upeps{K}) - \varepsilon q V(\upz{Z}_T \upz{K})\eta = A_{\varepsilon}+B_{\varepsilon}, \end{split} \end{equation} where \begin{equation} \label{equ:A-B} \begin{split} A_{\varepsilon}&:=\int_0^\varepsilon q \Big( V(\updel{Z}_T \upeps{K}) - V(\upz{Z}_T \upz{K}) \Big)\eta\, d\delta, \text{ and } \\ B_{\varepsilon}&:= \int_0^{\varepsilon}q V(\updel{Z}_T\upeps{K}) \Lambda\, \delta\, d\delta. \end{split} \end{equation} H\" older's inequality implies that \begin{equation} \label{equ:Holder} \begin{split} {\mathbb E}[ B_{\varepsilon}] \leq \tot \varepsilon^2 \sup_{\delta\in [0,\varepsilon]} \Big({\mathbb E}[ \updel{Z}_T \upeps{K}]^{-q} {\mathbb E}[ \Lambda^{1-p}]^{1+q}\Big) \leq \tot \varepsilon^2 {\mathbb E}[ \Lambda^{1-p}]^{1+q}. \end{split} \end{equation} Thus, we have $\tfrac{1}{\varepsilon} {\mathbb E}[ B_{\varepsilon}] \to 0$, as $\varepsilon\searrow 0$. To show that $\tfrac{1}{\varepsilon} {\mathbb E}[ A_{\varepsilon}]\to 0$, we note that ${\mathbb E}[ A_{\varepsilon}] = \int_0^{\varepsilon} f(\varepsilon,\delta)\, d\delta$, where the function $f:[0,\infty)^2 \to \R$ is given by \begin{equation} \label{equ:function-f} \begin{split} f(\varepsilon,\delta) := q\ee{ \big(V(\updel{Z}_T\upeps{K}) - V(\upz{Z}_T \upz{K}) \big)\eta}. \end{split} \end{equation} Since $f(0,0)=0$, it will be enough to show that $f$ is continuous at $(0,0)$. By the assumptions of the lemma and the definition of $\updel{Z}$, we have \[ V(\uppar{Z}{\delta_n}_T \uppar{K}{\varepsilon_n}) \to V(\upz{Z}_T \upz{K}),\text{ in probability},\] for each sequence $(\varepsilon_n,\delta_n)\in [0,\infty)^2$ such $(\varepsilon_n,\delta_n)\to (0,0)$. Therefore, it suffices to establish uniform integrability of the expression inside of the expectation in \eqref{equ:function-f}. For that we can use the theorem of de la Valle\' e-Poussin, whose conditions hold thanks to an application H\"older's inequality as in \eqref{equ:Holder} above, remembering that not only $\eta\in\el^{1-p}$, but also in $\el^s$, for some $s>(1-p)$. \end{proof} \begin{proof}[Proof of Theorem \ref{thm:main1}] Thanks to the optimality of $\upeps{Z}_T \upeps{\hat{H}}_T$, we have the upper estimate \begin{equation} \label{equ:upper} \begin{split} \oo{\varepsilon}\ee{ V(\upeps{Z}_T \upeps{\hat{H}}_T) - V(\upz{Z}_T \upz{\hat{H}}_T)} &\leq \oo{\varepsilon}\ee{ V(\upeps{Z}_T \upz{\hat{H}}_T) - V(\upz{Z}_T \upz{\hat{H}}_T)} \end{split} \end{equation} Similarly, we obtain the lower estimate \begin{equation} \label{equ:lower} \begin{split} \oo{\varepsilon}\ee{ V(\upeps{Z}_T \upeps{\hat{H}}_T) - V(\upz{Z}_T \upz{\hat{H}}_T)} &\geq \oo{\varepsilon}\ee{ V(\upeps{Z}_T \upeps{\hat{H}}_T) - V(\upz{Z}_T \upeps{\hat{H}}_T)}. \end{split} \end{equation} Our next task is to prove that the limits of the right-hand sides of \eqref{equ:upper} and \eqref{equ:lower} exist and both coincide with the right-hand side of \eqref{equ:main1-v}. In each case, Lemma \ref{lem:K-eps} can be applied; in the first with $\upeps{K}= \upz{\hat{H}}_T$, and in the second with $\upeps{K}=\upeps{\hat{H}}_T$. In both cases the assumption (1) of Lemma \ref{lem:K-eps} follows directly from that fact that $\upeps{Z}_T \upeps{K} \in \upeps{\sY}$. As for the assumption (2), it trivially holds in the first case. In the second case, we need to argue that $\upeps{\hat{H}}_T \to \upz{\hat{H}}_T$ in probability, as $\varepsilon\searrow 0$. That, in turn, follows easily from Lemma 3.10 in \cite{LarZit07}; as mentioned above, the seemingly stronger assumption of NFLVR made in \cite{LarZit07} is not necessary and its results hold under the weaker condition NUBPR. Having proven \eqref{equ:main1-v}, we turn to \eqref{equ:main1-u}. Thanks to \eqref{equ:simpl}, the conjugacy relationship \eqref{equ:conj} takes the following, simple, form in our setting: \begin{equation} \label{equ:conj2} \begin{split} p \upeps{u} = (q\upeps{v})^{1-p}. \end{split} \end{equation} Therefore, $\upeps{u}$ is right differentiable at $\varepsilon=0$, and we have \begin{equation*} \label{equ:6630} \begin{split} p \derep{\upeps{u}} &= (1-p) (q\upz{v})^{-p}\, q^2 \upz{v} \upz{\Delta} = p^2 \upz{u} \upz{\Delta}. \qedhere \end{split} \end{equation*} \end{proof} \subsection{Remaining proofs} \begin{proposition} Suppose that $\eta\in\el^{2(1-p)}$ and $\Lambda, \Lambda\eta\in \el^{1-p}$. Then \label{pro:v-up} for all $\varepsilon\geq 0$ we have \begin{equation} \label{equ:ve-up} \begin{split} \upeps{v} - \upz{v} - \varepsilon q \upz{v} \upz{\Delta} \leq \tot C_v \varepsilon^2 + \tot C'_v \varepsilon^3, \end{split} \end{equation} where $C_v= |q|\|\eta\|_{\el^{2(1-p)}}^{1/2} + \|\Lambda\|_{\el^{1-p}}$ and $C'_v =|q|\|\eta \Lambda\|_{\el^{1-p}}$. \end{proposition} \begin{proof} The upper estimate \eqref{equ:upper} and the representation \eqref{equ:primal1} imply that \[ \ee{ V(\upeps{Z}_T \upeps{\hat{H}}_T) - V(\upz{Z}_T \upz{\hat{H}}_T) - \varepsilon q V(\upz{Z}_T \upz{\hat{H}}_T)\eta} \leq {\mathbb E}[A_{\varepsilon}] + {\mathbb E}[B_{\varepsilon}],\] where $A_{\varepsilon}$ and $B_{\varepsilon}$ are defined by \eqref{equ:A-B}, with $\upeps{K}=\upz{K}= \upz{\hat{H}}_T$. As in \eqref{equ:Holder}, we have \[ {\mathbb E}[ B_{\varepsilon}]\leq \tot \varepsilon^2 \|\Lambda\|_{\el^{1-p}}.\] To deal with $A_{\varepsilon}$ we note that its structure allows us to apply the representation from \eqref{equ:primal1} once again to see \[ \oo{q^2} A_{\varepsilon} = \int_0^{\varepsilon} \int_0^{\delta} V(\uppar{Z}{\beta}_T \upz{\hat{H}}_T) \eta (\eta+\beta \Lambda)\, d\beta\, d\delta. \] This, in turn, can be estimated, via H\" older inequality, as in \eqref{equ:Holder}, as follows \[ {\mathbb E}[ A_{\varepsilon}] \leq \tot |q| \varepsilon^2 \sup_{ \beta\in [0,\varepsilon] } {\mathbb E}[ (\eta (\eta+\beta \Lambda))^{1-p}]^{1+q} \leq \tot |q| \varepsilon^2 \Big( \|\eta^2\|_{\el^{1-p}}+\varepsilon \|\eta \Lambda\|_{\el^{1-p}}\Big), \] yielding the bound in \eqref{equ:ve-up}. \end{proof} Unfortunately, the same idea cannot be applied to obtain a similar lower bound. Instead, we turn to the primal problem and establish a lower bound for it. \begin{proposition} \label{pro:u-down} Given $\varepsilon_0>0$, assume that $\Lambda\in\el^{1-p}$, and $\Phi^2 e^{ \varepsilon_0\abs{p} \Phi^-}\in\el^1(\upz{\tilde{\PP}})$, where $\upz{\tilde{\PP}}$ is defined by \eqref{equ:Girsanov}. Then, \[ \upeps{u} - \upz{u} - \varepsilon p \upz{u} \upz{\Delta} \geq - C_u(\varepsilon) \varepsilon^2 \text{ for } \varepsilon \in [0,\varepsilon_{0}],\] where $C_u(\varepsilon) := \tot p^2 |\upz{u}| {\mathbb E}^{\upz{\tilde{\PP}}}[ \Phi^2 e^{\varepsilon \abs{p} \Phi^-}]$. \end{proposition} \begin{proof} For $\tX := {\mathcal E}( \int_0^{\cdot} \upz{\hat{\pi}}_t\, d\Re_t)$, we have $\tX\in\upeps{\sX}$ so that, by optimality, \begin{equation} \label{equ:first-u} \begin{split} \upeps{u} - \upz{u} - p\varepsilon {\mathbb E}[ U(\upz{\hat{X}}_T) \Phi] &\geq {\mathbb E}[ U(\tX_T) - U(\upz{\hat{X}}_T) - p \varepsilon U(\upz{\hat{X}}_T)\Phi]. \end{split} \end{equation} Thanks to the form of $\tX$, the right-hand side of \eqref{equ:first-u} above can be written as ${\mathbb E}[ U(\upz{\hat{X}}_T) D_{\varepsilon}]$, where $D_{\varepsilon} = \exp( p \varepsilon \Phi) -1 - p\varepsilon \Phi = \int_0^{\varepsilon} \int_0^{\delta} p^2 \Phi^2 e^{p\beta \Phi} \, d\beta\, d\delta$. Thus, \begin{align}\label{lower_bound_est} \begin{split} {\mathbb E}[ U(\upz{\hat{X}}_T) D_{\varepsilon}] &= p^2 \int_0^{\varepsilon} \int_0^{\delta} {\mathbb E}[ U(\upz{\hat{X}}_T) \Phi^2 e^{ p \beta \Phi}]\, d\beta\, d\delta \\&\geq \tot p^2 \varepsilon^2{\mathbb E}[ U(\upz{\hat{X}}_T) \Phi^2 e^{ \varepsilon \abs{p} \Phi^-}]. \end{split} \end{align} Therefore, $\upeps{u} - \upz{u} - \varepsilon p {\mathbb E}[ U(\upz{\hat{X}}_T) \Phi] \geq - C_u(\varepsilon) \varepsilon^2$, for $\varepsilon\in [0,\varepsilon_0]$ with $C_u$ as in the statement. It remains to show that ${\mathbb E}[ U(\upz{\hat{X}}_T) \Phi] = {\mathbb E}[ U(\upz{\hat{X}}_T) \eta]$ which is equivalent to showing ${\mathbb E}^{\tilde{\mathbb P}^{(0)}}[ \Phi] = {\mathbb E}^{\tilde{\mathbb P}^{(0)}}[\eta]$ by the definition of $\tilde{\mathbb P}^{(0)}$. We define the local $\tilde{\mathbb P}^{(0)}$-martingale $\tM^p$ by \eqref{equ:tMp}. Therefore, $N = \int_0^{\cdot} \lambda'_t\, d\tM^p_t$ is also a local martingale. The desired equality is therefore equivalent to the equality ${\mathbb E}^{\upz{\tilde{\PP}}}[N_T]=0$ by the definition of $\eta$ and $\Phi$. In turn, it is sufficient to show that $N$ is an $\sH^2$-martingale under $\upz{\tilde{\PP}}$. Since $\ab{N}_T = \int_0^{T} (\lambda'_t)^2\, d\ab{M}_t = \Lambda$ , H\" older's inequality implies that \begin{equation*} \label{equ:49B0} \begin{split} {\mathbb E}^{\tilde{\mathbb P}^{(0)}}[\ab{N}_T] & = (q\upz{v})^{-1} {\mathbb E}[(\upz{\hat{Y}}_T)^{-q} \Lambda] \leq (q\upz{v})^{-1} {\mathbb E}[\Lambda^{1-p}]^{1+q}<\infty. \qedhere \end{split} \end{equation*} \end{proof} \revised{ \begin{remark} If one is interested in an error estimate which does not feature the optimal portfolio $\upz{\hat{\pi}}$ (through $\Phi$), one can adopt an alternative approach in the proof (and the statement) of Proposition \ref{pro:u-down}. More specifically, by using $\tX = \upz{\hat{X}}{\mathcal E}( \int_0^{\cdot} \varepsilon \lambda' \, d\Re_t)$ as a test process (instead of ${\mathcal E}( \int_0^{\cdot} \upz{\hat{\pi}}_t\, d\Re_t)$), one obtains a constant $C_u(\varepsilon)$ which depends only on the primal and dual optimizers $\upz{\hat{X}}$ and $\upz{\hat{Y}}$, in addition to $\lambda'$, $\eta$ and $\Lambda$. \end{remark} } \begin{proof}[Proof of Theorem \ref{thm:main2}] Two of the four inequalities in Theorem \ref{thm:main2} have been established in Propositions \ref{pro:v-up} and \ref{pro:u-down}. For the remaining two we use the special form \eqref{equ:conj2} of the conjugacy relationship between $\upeps{u}$ and $\upeps{v}$. Thanks to Proposition \ref{pro:u-down} and the positivity of $p\upeps{u}$, $q\upeps{v}$ and $1+q$, we have \begin{equation*} \label{equ:4797} \begin{split} q\Big( \upeps{v}- \upz{v} - \varepsilon q \upz{v} \upz{\Delta}\Big) &= (p\upeps{u})^{1+q} - (p\upz{u})^{1+q} - \varepsilon q (p\upz{u})^{1+q}\upz{\Delta}. \end{split} \end{equation*} The right-hand side above is further bounded from above, for $\varepsilon$ in a (right) neighborhood of $0$, by \[ F(\varepsilon) := (p\upz{u} + \varepsilon p\upz{u} \upz{\Delta} - p C\varepsilon^2 )^{1+q} - (p\upz{u})^{1+q} - \varepsilon q (p\upz{u})^{1+q} \upz{\Delta},\] where $C$ is the constant from Proposition \ref{pro:u-down}. $F$ is a $C^2$-function in some neighborhood of $0$ with $F(0)=F'(0)=0$; hence, on each compact subset of that neighborhood it is bounded by a constant multiple of $\varepsilon^2$. In particular, we have \[ \upeps{v} - \upz{v} - \varepsilon q \upz{v} \upz{\Delta} \geq - C \varepsilon^2,\] for some $C>0$ and $\varepsilon$ in some (right) neighborhood of $0$. A similar argument, but based on Proposition \ref{pro:v-up}, shows that \eqref{equ:main2-u} holds, as well. \end{proof} \proof[Proof of Theorem \ref{thm:main3}] The first part of \eqref{2nd_int} means that $\Lambda \in \el^{1-p}({\mathbb P})$; hence, the second half of the proof of Proposition \ref{pro:u-down} shows that ${\mathbb E}^{\upz{\tilde{\PP}}}[\Phi] = \Delta^{(0)}$. Therefore, the martingale representation \eqref{Itorep} can be written as \begin{align}\label{Itorep1} \Phi = \Delta^{(0)} + \int_0^T \gamma^B_{t}\sigma_tdB^{\upz{\tilde{\PP}}}_t + \int_0^T \gamma^W_{t} dW^{\upz{\tilde{\PP}}}_t. \end{align} Because the filtration is generated by the Brownian motions $(B,W)$ we can find $\hat{\nu}^{(0)} \in \sP_{W}^2$ such that the dual optimizer $\hat{H}^{(0)}$ can be represented as $$ \hat{H}^{(0)} = \mathcal{E}(-\int \hat{\nu}^{(0)} dW). $$ Therefore, Girsanov's Theorem ensures that under $\upz{\tilde{\PP}}$, the processes $$ dB^{\upz{\tilde{\PP}}} := dB +(\lambda -\hat{\pi}^{(0)})\sigma dt, \quad \text{ and } \quad dW^{\upz{\tilde{\PP}}}:= dW +\hat{\nu}^{(0)}dt, $$ are independent Brownian motions. We start with the primal problem and define $\tilde{\pi} := \hat{\pi}^{(0)} + \varepsilon \delta$ with $\delta:=q\gamma^B +\frac{\lambda'}{1-p}\in \sP_{B}^2$. Then we have \begin{align*} (\tilde{X})^{p} &:=\mathcal{E}\big(\int\tilde{\pi}dR^{(\varepsilon)})^p\\ &= \big(\hat{X}^{(0)}\big)^pe^{p\int\big (\varepsilon\hat{\pi}^{(0)} \lambda'+ \varepsilon^2 (\delta \lambda'-\frac12\delta^2 ) \big)\sigma^2dt + p\varepsilon \int\delta\sigma dB^{\upz{\tilde{\PP}}} }. \end{align*} Consequently, by replacing $e^x$ with its Taylor expansion and using that the involved $\upz{\tilde{\PP}}$-expectation is finite (here we use the integrability requirement \ref{exp_reg}), we find a function $C_u(\varepsilon) \in O(\varepsilon^3)$ such that \begin{align}\label{Cu} \begin{split} {\mathbb E}[U(\tilde{X}_T)] &= u^{(0)} {\mathbb E}^{\upz{\tilde{\PP}}}\left[e^{p\int_0^T \big(\varepsilon\hat{\pi}^{(0)} \lambda'+ \varepsilon^2 (\delta \lambda'-\frac12\delta^2)\big)\sigma^2dt +p\varepsilon\int_0^T \delta\sigma dB^{\upz{\tilde{\PP}}} }\right]\\ & = u^{(0)}\Big(1+p\varepsilon\Delta^{(0)}+ \frac12 p\varepsilon^2\Big\{p(\Delta^{(0)})^2 +\Delta^{(00)}\Big\}\Big)+ C_u(\varepsilon). \end{split} \end{align} We then turn to the dual problem. For the perturbed dual control $\tilde{\nu}:= \hat{\nu}^{(0)} -\varepsilon p\gamma^W\in\sP_{W}^2$ we have \begin{align*} &\Big(Z^{(\varepsilon)} \mathcal{E}(-\int \tilde{\nu} dW)\Big)^{-q}\\&= e^{q\int ( \lambda+\varepsilon\lambda' )\sigma dB+q\int ( \hat{\nu}^{(0)}-\varepsilon p\gamma^W )dW+q\frac12\int \big(( \lambda+\varepsilon \lambda' )\sigma ^2 + | \hat{\nu}^{(0)}-\varepsilon q\gamma^W |^2 \big)dt}\\ &= (Z^{(0)}\hat{H}^{(0)})^{-q}e^{\varepsilon q\int \lambda'\sigma dB^{\upz{\tilde{\PP}}}-\varepsilon qp\int \gamma^W dW^{\upz{\tilde{\PP}}}+q\frac12\int \big( \varepsilon^2 (\lambda')^2\sigma^2 +\varepsilon^2 p^2|\gamma^W|^2+2\varepsilon\lambda'\pi^{(0)}\sigma^2\big)dt}. \end{align*} Since $\tilde{\nu}$ is admissible in the $\varepsilon$-problem we find \begin{align* &v^{(\varepsilon)} \le \frac1q{\mathbb E}\left[\Big(Z_T^{(\varepsilon)} \mathcal{E}(-\int_0^T \tilde{\nu} dW)\Big)^{-q}\right] \\&= v^{(0)} {\mathbb E}^{\upz{\tilde{\PP}}}\left[e^{\varepsilon q\int_0^T \lambda'\sigma dB^{\upz{\tilde{\PP}}}-\varepsilon qp\int_0^T \gamma^W dW^{\upz{\tilde{\PP}}}+q\frac12\int_0^T \big( \varepsilon^2 (\lambda')^2\sigma^2 +\varepsilon^2p^2 |\gamma^W|^2+2\varepsilon\lambda'\pi^{(0)}\sigma^2\big)dt}\right]. \end{align*} Finiteness of $v^{(\varepsilon)}$ ensures that the $\upz{\tilde{\PP}}$-expectation appearing on the last line is also finite (recall that $q<0$). As in the primal problem, this allows us to replace $e^x$ with its Taylor series and in turn implies that we can find a function $C_v(\varepsilon) \in O(\varepsilon^3)$ such that \begin{align*} &v^{(0)} {\mathbb E}^{\upz{\tilde{\PP}}}\left[e^{\varepsilon q\int_0^T \lambda'\sigma dB^{\upz{\tilde{\PP}}}-\varepsilon qp\int_0^T \gamma^W dW^{\upz{\tilde{\PP}}}+q\frac12\int_0^T \big( \varepsilon^2 (\lambda')^2\sigma^2 +\varepsilon^2p^2 |\gamma^W|^2+2\varepsilon\lambda'\pi^{(0)}\sigma^2\big)dt}\right]\\&= v^{(0)}\Big(1+q\varepsilon\Delta^{(0)} + \frac12q\varepsilon^2\Big\{q(\Delta^{(0)})^2 + \Delta^{(00)}\Big\}\Big)+ C_v(\varepsilon). \end{align*} By combining this estimate and \eqref{Cu} with the primal-dual relation \eqref{equ:conj2} we find \begin{align} \label{fenchel} \begin{split} & u^{(0)}\Big(1+p\varepsilon\Delta^{(0)}+ \frac12 p\varepsilon^2\Big\{p(\Delta^{(0)})^2 +\Delta^{(00)}\Big\}\Big)+ C_u(\varepsilon) \\&\le u^{(\epsilon)} \\&=\frac1p (q v^{(\epsilon)})^{1-p}\\& \le \frac1p\Big(q v^{(0)}\Big[1+q\epsilon \Delta^{(0)} + \frac12q\epsilon^2\Big\{q(\Delta^{(0)})^2 + \Delta^{(00)}\Big\}+C_v(\epsilon)\Big]\Big)^{1-p}. \end{split} \end{align} The function $x\to x^{1-p}$ is real analytic on $(0,\infty)$. Therefore, the fact that $C_v\in O(\varepsilon^3)$ ensures that the last line of \eqref{fenchel} agrees with the first line of \eqref{fenchel} up to $O(\varepsilon^3)$-terms. This establishes \eqref{equ:main3-u}. A similar argument produces \eqref{equ:main3-v}. \endproof \section{Examples} \label{sec:examples} \subsection{First examples} We start this section with a short list of trivial and extreme cases. They are not here to illustrate the power of our main results, but simply to help the reader understand them better. They also tell a similar, qualitative, story: loosely speaking, the improvement in the utility (on the log scale) is proportional both to the base market price of risk process and to the size of the deviation. Locally, around $\lambda$, the value function of the utility maximization problem - parametrized by the market price of risk process $\tilde{\lambda}$ - is well approximated by an exponential function of the form \begin{equation} \label{equ:exp-approx} \begin{split} u(\tilde{\lambda}) \approx u(\lambda) e^{\scl{\tilde{\lambda}-\lambda}{\upz{\hat{\pi}}}_{0}},\text{ where } \scl{\rho}{\pi}_{0} = {\mathbb E}^{\upz{\tilde{\PP}}}[ \int_0^T \rho_t\pi_t\, dt], \end{split} \end{equation} where $u(\tilde{\lambda})$ and $u(\lambda)$ denote the values of the utility-maximization problems with market price of risk processes $\tilde{\lambda}$ and $\lambda$, respectively. \begin{example}[Small market price of risk] \label{exa:small} Suppose that $\lambda\equiv 0$ so that we can think of $\upeps{S}$ as the stock price in a market with a ``small'' market price of risk. Since $\upz{Z}\equiv 1$, it is clearly the dual optimizer at $\varepsilon=0$ and we have $\upz{\hat{\pi}}\equiv 0$. Consequently, under the assumptions of Theorem \ref{thm:main2}, we have $\upz{\tilde{\PP}}={\mathbb P}$ and \[ \upz{\Delta} = {\mathbb E}^{\upz{\tilde{\PP}}}[ \int_0^T \lambda'_t\, dM_t]=0.\] It follows that \[ \upeps{u} = \upz{u} + O(\varepsilon^2)\text{ and } \upeps{v}=\upz{v} + O(\varepsilon^2),\] and the effects of $\varepsilon \lambda'$ are felt only in the second order, regardless of the risk-aversion coefficient $p<0$. \end{example} \begin{example}[Deviations from the Black-Scholes model] \label{exa:Black-Scholes} Suppose that $M=B$ is an ${\mathbb F}$-Brownian motion and that $\lambda\ne 0$ is a constant process (we also use $\lambda$ for the value of the constant). In that case, it is classical that the dual minimizer in the base market is $\upz{Z} = {\mathcal E}( - \lambda B)$ and, consequently, that $\rn{\upz{\tilde{\PP}}}{{\mathbb P}} = {\mathcal E}( q\lambda B)$. It follows that \[ \upz{\Delta} = \tfrac{\lambda}{1-p} {\mathbb E}^{\upz{\tilde{\PP}}}[ \int_0^T \lambda'_t\, dt].\] As we will see below, this form is especially convenient for computations. \end{example} \begin{example}[Uniform deviations] \label{exa:constant} Another special case where it is particularly easy to compute the (logarithmic derivative) $\upz{\Delta}$ is when the perturbation $\lambda'$ is a constant process (whose value is also denoted by $\lambda'$). Indeed, in that case \begin{equation} \label{equ:Dz-unif} \begin{split} \upz{\Delta} = \lambda' {\mathbb E}^{\upz{\tilde{\PP}}}[ \int_0^T \upz{\hat{\pi}}_t\, dt]. \end{split} \end{equation} It is especially instructive to consider the case where the base model is Black and Scholes' model since everything becomes explicit: the optimal portfolio is given by the Merton proportion $\upz{\hat{\pi}}_t = \lambda/(1-p)$, and the the values $\upeps{u}$ and $\upeps{v}$ are given by \[ p\upz{u} = \exp(\tot q \lambda^2 T) \text{ and } q\upz{v} = \exp(\tot \tfrac{q}{1-p} \lambda^2 T). \] Using \eqref{equ:Dz-unif} or by performing a straightforward direct computation, we easily get \[ p\upz{\Delta} = q \lambda' \lambda T, \] making the approximation in \eqref{equ:exp-approx} exact. \end{example} \subsection{The Kim-Omberg model} The Kim-Omberg model (see \cite{KimOmb96}) is one of the most widely used models for the market price of risk process. Because the Kim-Omberg model allows for explicit expressions for all quantities involved in CRRA utility maximization it serves as an excellent test case for the practical implementation of our main results. We assume that ${\mathbb F}$ is the augmentation of the filtration generated by two independent one dimensional Brownian motions $B$ and $W$ and define $\lambda^\text{KO}$ be the Ornstein-Uhlenbeck process \begin{align} d\lambda_t^\text{KO} &:= \kappa (\theta - \lambda_t^\text{KO})dt + \beta dB_t + \gamma dW_t, \quad \lambda^\text{KO}_0\in\R\label{OU}, \end{align} where $\kappa, \theta, \beta$ and $\gamma$ are constants. We define the volatility $M_t := B_t$ in what follows. The following result summarizes the main properties in \cite{KimOmb96}: \begin{theorem}[Kim and Omberg 1996]\label{thm:KO} Let the market price of risk process be defined by \eqref{OU}, $M:=B$, and let $p<0$. Then there exist continuously differentiable functions $a,b,c:[0,\infty)\to \R$ such that for $t\in [0,T)$ we have \begin{align*} -a'(t) & = \alpha_1\, b(t) + \tot \alpha_3\, c(t) - \tot \alpha_2\, b^2(t), & a(T)&=0, \\ -b'(t) &= \alpha_4\, b(t)+ \alpha_1\, c(t) - \alpha_2\, b(t) c(t),& b(T)&=0, \\ -c'(t) &= -q + 2 \alpha_4\, c(t) - \alpha_2\, c^2(t), & c(T)&=0, \end{align*} where $\alpha_1:=\theta\kappa$, $\alpha_2 := (1+q) \beta^2+ \gamma^2$, $\alpha_3 := \beta^2+\gamma^2$ and $\alpha_4 := q \beta - \kappa$. Furthermore, the primal value function reads \begin{align} \label{equ:exact} u^\text{KO}(x) = \frac{x^p}{p}e^{-a(0) - b(0)\lambda^\text{KO}_0 - \tfrac12 c(T)(\lambda^\text{KO}_0)^2},\quad x>0, \end{align} and the corresponding primal optimizer is given by \begin{align}\label{KO_primal} \hat{\pi}^\text{KO}_t = \frac{b(t)\beta + \big(c(t)\beta-1\big) \lambda^\text{KO}_t}{p-1},\quad t\in[0,T]. \end{align} \end{theorem} For $p<0$, the above Riccati equation describing $c$ has the ``normal non-exploding solution'' as defined in the appendix of \cite{KimOmb96}. Therefore, all three functions $a,b$, and $c$ are bounded on any finite time-interval $[0,T]$ of $(0,\infty)$. To illustrate our approximation we think of the Kim-Omberg model as a perturbation of a base model. As base model we will consider the following model with ``totally-unhedgable-coefficients'' (see Example 7.4, p.~305, in \cite{KarShr98}): \begin{align}\label{KO1} d\lambda_t := \kappa (\theta-\lambda_t) dt + \gamma \, dW_t,\quad \lambda_0:=\lambda^\text{KO}_0. \end{align} This way, $\ld^\text{KO} = \lambda+\varepsilon \lambda'$, where $\varepsilon = \beta$ and \begin{align}\label{KO2} d\lambda'_t := - \kappa \lambda'_t\, dt + dB_t,\quad \lambda'_0:=0. \end{align} The following result provides closed-form expressions for our correction terms: \begin{lemma}\label{lem:KO} Let $(\lambda,\lambda')$ be defined by \eqref{KO1}-\eqref{KO2} and let $p<0$. For the $\varepsilon =0$ model the primal and dual optimizers are given by \begin{align}\label{KO:base} \hat{\pi}^{(0)}_t = \frac{\lambda_t}{1-p},\quad \hat{\nu}^{(0)}_t = \gamma \Big(b(t)+c(t)\lambda_t\Big),\quad t\in[0,T]. \end{align} Furthermore, the processes $(\gamma^B,\gamma^W)$ appearing in the martingale representation \eqref{Itorep} of $\Phi$ are given by \begin{align} \gamma_t^B &=\tfrac1{p-1}(C_2(t)+C_6(t)\lambda_t),\label{KOgammaB}\\ \gamma^W_t &= \tfrac{\gamma}{p-1}(C_4(t)+2C_5(t) \lambda_t+C_6(t)\lambda_t'),\label{KOgammaW} \end{align} where the functions $C_1,C_2,C_4, C_5$ and $C_6$ in \eqref{KOgammaB}-\eqref{KOgammaW} satisfy the ODEs \begin{align*} -C_1'(t)&= \tilde{b}(t) \, C_4(t)+ \gamma^2 \, C_5(t),& C_1(T)&=0,\\ -C_2'(t)&= \tilde{b}(t) \, C_6(t) - \kappa \, C_2(t),& C_2(T)&=0,\\ -C_4'(t) &= q\, C_2(t) - \tilde{c}(t) \, C_4(t) + 2\tilde{b}(t) \, C_5(t), & C_4(T)&=0,\\ -C_5'(t) &= q\, C_6(t) -2 \tilde{c}(t) \, C_5(t),& C_5(T)&=0,\\ -C'_6(t) &= - (\kappa +\tilde{c}(t)) \, C_6(t) -1,& C_6(T)&=0, \end{align*} on $[0,T)$, with $(a,b,c)$ as in Theorem \ref{thm:KO} (with $\beta:=0$), $\tilde{b}(t):=\kappa\theta-\gamma^2 b(t)$ and $\tilde{c}(t) := \kappa+\gamma^2 c(t) $. Furthermore, for the measure $\tilde{\mathbb P}^{(0)}$ defined by \eqref{equ:Girsanov} and for all $T>0$ we have \begin{align}\label{Delta0_KO} \Delta^{(0)} &:= {\mathbb E}^{\tilde{\mathbb P}^{(0)}}\left[ \int_0^T \lambda'_s\hat{\pi}_s^{(0)} ds\right]= -\frac1{1-p}\Big(C_1(T) + C_4(T) \lambda_0 +C_5(T) \lambda_0^2\Big), \end{align} \end{lemma} \proof The first part follows from Theorem \ref{thm:KO} applied to the case $\beta:=0$. To find the martingale representation \eqref{Itorep} we define the function $$ f(t,x,\lambda):= \frac{x^p}{p}e^{-a(t) - b(t)\lambda - \tfrac12 c(t)\lambda^2},\quad t\in[0,T],\quad x>0,\quad \lambda\in\R, $$ where the functions $(a,b,c)$ are as in Theorem \ref{thm:KO}. The martingale properties of $f(t,\hat{X}^{(0)}_t,\lambda_t)$ and $\hat{X}^{(0)}_t\hat{Y}^{(0)}_t$ as well as the proportionality property $(\hat{X}^{(0)}_T)^{p}\propto\hat{X}^{(0)}_T \hat{Y}^{(0)}_T$ produce $$ pf(t, \hat{X}^{(0)}_t,\lambda_t)=p{\mathbb E}[f(T, \hat{X}^{(0)}_T,\lambda_T)|\sF_t]\propto{\mathbb E}[ \hat{Y}^{(0)}_T\hat{X}^{(0)}_T|\sF_t] = \hat{X}^{(0)}_t\hat{Y}^{(0)}_t. $$ By computing the dynamics of the left-hand-side we see from Girsanov's Theorem that the two processes \begin{align*} dB^{\tilde{\mathbb P}^{(0)}}_t &:= - q\lambda_tdt + dB_t,\\ \quad dW^{\tilde{\mathbb P}^{(0)}}_t &:= \Big(b(t)+c(t)\lambda_t\Big)\gamma dt + dW_t, \end{align*} are independent Brownian motions under $\tilde{\mathbb P}^{(0)}$. These dynamics and It\^o's Lemma ensure that \begin{align*} N_t := \int_0^t &\lambda'_s\lambda_s ds -C_1(t) -C_2(t)\lambda'_ - C_4(t) \lambda_t -C_5(t) \lambda_t^2-C_6(t)\lambda_t\lambda'_t, \end{align*} is a $\tilde{\mathbb P}^{(0)}$-local martingale. Because the processes $(\lambda,\lambda')$ remain Ornstein-Uhlenbeck processes under $\tilde{\mathbb P}^{(0)}$ and the functions $C_1$-$C_6$ are bounded, $N$ is indeed a $\tilde{\mathbb P}^{(0)}$-martingale. Furthermore, thanks to the zero terminal conditions imposed on $C_1$-$C_6$, we see that \begin{align*} \Phi=\tfrac1{1-p}\int_0^T \lambda_t \lambda'_tdt &= \tfrac1{1-p}N_T = \tfrac1{1-p}N_0 +\int_0^T \gamma^B_t dB^{\upz{\tilde{\PP}}}_t +\int_0^T \gamma^W_t dW^{\upz{\tilde{\PP}}}_t, \end{align*} for $(\gamma^B,\gamma^W)$ defined by \eqref{KOgammaB}-\eqref{KOgammaW}. \endproof \subsubsection{Exact computations.} The proof of Lemma \ref{lem:KO} shows that $$ \Delta^{(0)} :={\mathbb E}^{\tilde{\mathbb P}^{(0)}}\left[ \int_0^T \lambda'_s\hat{\pi}_s^{(0)} ds\right]= \frac1{p-1}\Big(C_1(0) + C_4(0) \lambda_0 +C_5(0) \lambda_0^2\Big). $$ This relation, a similar one (whose exact form and the derivation we omit) for the second-order term $\Delta^{(00)}$ of \eqref{Delta00}, and the availability of the exact expression \eqref{equ:exact} for the value function $u^{KO}$ allow for an efficient numerical computation of the zeroth-, first-, and second-order approximation, and their comparison with the exact values. The model parameters used in the below Table 1 are the calibrated model parameters for the market portfolio reported in Section 4.2 in \cite{LarMun2012} (we ignore the constant interest rate and constant volatility used in Section 4 in \cite{LarMun2012}). Moreover, we use negative values of $\varepsilon$ because the empirical covariation between excess return and the stock's return is typically negative (see, e.g., the discussion in Section 4.2 in \cite{LarMun2012}). Instead of hard-to-interpret expected utility values, we report their certainty equivalents (i.e., their compositions with the function $CE:=U^{-1}$; see Remark \ref{rem:761C}(1)). We set $\delta^{(0)} := p u^{(0)} \Delta^{(0)}$ and $\delta^{(00)} :=p u^{(0)} \big( \Delta^{(00)} + p (\Delta^{(0)})^2\big)$. \begin{center} \begin{tabular}{cc||ccc|c} $\varepsilon$ & $\lambda_0$& $CE(u^{(0)}) $& $CE(u^{(0)}+ \varepsilon \delta^{(0)})$& $CE(u^{(0)}+ \varepsilon \delta^{(0)} + \tfrac{\varepsilon^2}{2} \delta^{(00)})$& $CE(u^{(\varepsilon)})$\\ \hline\hline -0.01 & 0.1 & 1.046 & 1.047 & 1.048 & 1.048 \\ - 0.05 & 0.1 & 1.046 & 1.054 & 1.081 & 1.084 \\ - 0.10 & 0.1 & 1.046 & 1.063 & 1.181 & 1.206 \\\hline - 0.01 & 0.5 & 1.614 & 1.647 & 1.648 & 1.649 \\ - 0.05 & 0.5 & 1.614 & 1.794 & 1.850 & 1.846 \\ - 0.10 & 0.5 & 1.614 & 2.020 & 2.339 & 2.272 \\ \hline \end{tabular} \footnotesize \begin{quotation} \textbf{Table 1.} Certainty equivalents for the zeroth-, first-, and second-order approximations and the exact values in the Kim-Omberg model with $\beta:=\varepsilon$ and unit initial wealth. The model parameters used are $\gamma :=0.04395,\,\kappa := 0.0404,\,\theta := 0.117,\,p:=-1$, and $T:=10$. \end{quotation} \end{center} \subsubsection{Monte-Carlo-based computations} One of the advantages of our approach is that it lends itself easily to computational methods based on Monte-Carlo (MC) simulation. For the Kim-Omberg model we use the standard explicit Euler scheme from MC simulation to compute the involved quantities of interest. In other words, we do not rely on the availability of exact expressions for the value functions or the correction terms $\Delta^{(0)}$ and $\Delta^{(00)}$. For a portfolio $\pi$ and the model-perturbation parameter $\varepsilon$, the constant $\text{CE}^{(\varepsilon)}(\pi)\in(0,\infty)$ is uniquely defined by \begin{align}\label{approxCE} U\left(\text{CE}^{(\varepsilon)}(\pi)\right) = {\mathbb E}\left[U\left({\mathcal E}\big(\int_0^T \pi_t\, d\Re_t\big)\right)\right]. \end{align} In other words, $\text{CE}^{(\varepsilon)}{(\pi)}$ is the dollar amount whose utility value matches that of the expected utility an investor would obtain in the $\varepsilon$-model who uses the strategy $\pi$. We remind the reader that $\hat{\pi}^{(0)}$ denotes the optimizer in the base ($\varepsilon=0$) model, $\tilde{\pi}^{(\varepsilon)}$ is the second-order improvement (as in \ref{optimal_control_correction} above) of $\hat{\pi}^{(0)}$, and $\hat{\pi}^{(\varepsilon)}$ is the exact optimizer in the $\varepsilon$-model. Both quantities $\text{CE}^{(\varepsilon)}(\hat{\pi}^{(0)})$ and $\text{CE}^{(\varepsilon)}(\tilde{\pi}^{(\varepsilon)})$ serve as lower bounds for the exact value $\text{CE}(u^{(\varepsilon)})$. The second one, which we also denote by \begin{align}\label{def_LB} \text{LB} := \text{CE}^{(\varepsilon)}(\tilde{\pi}^{(\varepsilon)}), \end{align} is second-order optimal and appears in our simulations. To obtain a corresponding upper bound, we simulate the dynamics of the dual process, based on $\eqref{equ:conj2}$ and the second-order optimal dual control $\tilde{\nu}$ defined by \eqref{optimal_dual_control_correction}. We define \begin{align}\label{def_UB} \text{UB} := U^{-1} \left( \frac1p{\mathbb E}\left[\Big(Z_T^{(\varepsilon)} \mathcal{E}(-\int_0^T \tilde{\nu}_u dW_u)\Big)^{-q}\right]^{1-p} \right). \end{align} To quantify the simulation errors, we report the $95\%$-confidence intervals based on MC simulated values of $\text{CE}^{(\varepsilon)}(\hat{\pi}^{(0)})$, LB, and UB in the below Table 2. The value $\text{CE}(u^{(\varepsilon)})$, computed without MC simulation and included for comparison only, is exact to $3$ decimal places. \begin{center} \begin{tabular}{cc||ccc|c} $\varepsilon$ & $\lambda_0$& $\text{CE}^{(\varepsilon)}(\hat{\pi}^{(0)}) $ & $\text{LB}$ & $\text{UB}$ & $\text{CE}(u^{(\varepsilon)})$\\ \hline\hline -0.01& 0.10& $[1.047, 1.048]$& [1.048, 1.049]& [1.048, 1.049]& 1.048\\ -0.05& 0.10& $[1.052, 1.053]$& [1.083, 1.084]& [1.083, 1.085]& 1.084\\ -0.10& 0.10& $[1.057, 1.058]$& [1.200, 1.201]& [1.204, 1.208]& 1.206\\ \hline -0.01& 0.50& $[1.644, 1.649]$& [1.647, 1.653]& [1.646, 1.657]& 1.649\\ -0.05& 0.50& $[1.760, 1.764]$& [1.844, 1.850]& [1.843, 1.857]& 1.846\\ -0.10& 0.50& $[1.868, 1.871]$& [2.248, 2.256]& [2.266, 2.286]& 2.272\\ \hline \end{tabular} \begin{quotation} \footnotesize \textbf{Table 2.} $95\%$-confidence intervals for certainty equivalents for the upper and lower bounds as well as the base model optimizer $\hat{\pi}^{(0)}$ for the Kim-Omberg model. The true exact values for the $\varepsilon$-model are included in the last column for comparison. Except for the last column, the numbers are based on MC simulation using Euler's scheme with one million paths each with time-step size $0.001$. The model parameters are the same as in Table 1. \end{quotation} \end{center} In Table 2 we note the significant difference between the performance of the base-model optimizer $\hat{\pi}^{(0)}$ and its second-order improvement $\tilde{\pi}^{(\varepsilon)}$; especially for larger values of $\varepsilon$. Furthermore, the lower and upper bounds appear to be quite tight. \subsection{Extended affine models} \label{sse:extaff} We turn to a class of models for which no closed-form expressions for the value functions $u$ and $v$ seem to be available. It constitutes the main example of the class of so-called extended-affine specifications of the market price of risk models introduced by \cite{CheFilKim07}. As in the Kim-Omberg model above we let the augmented filtration be generated by two independent Brownian motions $B$ and $W$. The central role is played by the following Feller process $F$ \begin{align} dF_t&:= \kappa (\theta - F_t)dt + \sqrt{F_t}\, \big( \beta dB_t + \gamma dW_t\big), \quad F_0>0,\label{Feller} \end{align} where $\kappa, \theta, \beta$ and $\gamma$ are strictly positive constants such that the (strict) Feller condition $2\kappa\theta > \beta^2 +\gamma^2$ holds. This ensures, in particular, that $F$ is strictly positive on $[0,T]$, almost surely. Unlike in the Kim-Omberg model, the appropriate volatility normalization turns out to be $\sqrt{F_t}$; that is, we define \begin{equation} \label{equ:M-F} \begin{split} M := \int_0^{\cdot} \sqrt{F_t}\, dB_t. \end{split} \end{equation} A particular extended affine specification of the market price of risk process considered in \cite{CheFilKim07} is given by \begin{align}\label{CFK} \lambda^\text{CFK}_t := \frac{\varepsilon}{F_t}+1, \end{align} where $\varepsilon$ is a (positive or negative) constant. Unless $\varepsilon=0$, there is currently no known closed-form solution to the corresponding optimal investment problem (Theorem 4.5 in \cite{GR15} expresses the corresponding value function as an infinite sum of weighted generalized Laguerre polynomials). However, for $\varepsilon =0$, the resulting model is covered by the analysis in \cite{Kra05}. Therefore, we choose the constant market price of risk process \[ \lambda_t:= 1\] for the base model whereas we define the perturbation process $\lambda'$ by \begin{align}\label{CV} \lambda'_t := \frac1{F_t}. \end{align} \begin{theorem}[Kraft 2005]\label{thm:CV} For $p<0$ there exist continuously differentiable functions $a,b:[0,T)\to \R$ such that \begin{align*} - a'(t)&= \alpha_1\, b(t),& a(T)&=0, \\ -b'(t) &= \alpha_4 \, b(t) - \tot \alpha_2 \, b^2(t) - \tot q,& b(T)&=0, \end{align*} where $\alpha_1:=\theta\kappa$, $\alpha_2 := (1+q) \beta^2+ \gamma^2$, and $\alpha_4 := q \beta - \kappa$. The value function of the utility-maximization problem with $\lambda:= 1$ and $M$ as in \eqref{equ:M-F} is given by $$ u^{(0)}(x) = \frac{x^p}{p}e^{-a(0) - b(0) F_0},\quad x>0. $$ The corresponding primal and dual optimizers are given by \begin{align}\label{CV_primal} \hat{\pi}^{(0)}_t = \frac{b(t)\beta-1}{p-1},\quad \hat{\nu}^{(0)}_t = b(t)\gamma\sqrt{F_t},\quad t\in[0,T]. \end{align} \end{theorem} To check the conditions of our main theorems, we use the explicit expression in \cite{HurKuz2008}, Theorem 3.1, for the Laplace transform \[ L(a_1, a_2) := {\mathbb E}[ \exp( a_1 Q + a_2 \Lambda)],\quad Q:=\int_0^T F_s\, ds,\quad \Lambda:=\int_0^T \tfrac{1}{F_s}\, ds.\] It is shown in \cite{HurKuz2008} that $L$ is finite in some neighborhood of $0$ under the strict Feller condition $2\kappa\theta > \beta^2 +\gamma^2$. This implies that both $\Lambda$ and $Q$ have a finite exponential moment. In particular, H\"older's inequality with exponents $-1/q$ and $(1-p)$ implies that $$ {\mathbb E}^{\upz{\tilde{\PP}}}[\Lambda]=\frac1{qv^{(0)}}{\mathbb E}[(\hat{Y}_T^{(0)})^{-q}\Lambda]\le \frac1{qv^{(0)}}{\mathbb E}[\Lambda^{1-p}]^{\frac{1}{1-p}}<\infty. $$ Thanks to the deterministic behavior of $\hat{\pi}^{(0)}$ in \eqref{CV_primal}, the martingale representation \eqref{Itorep} of $\Phi$ holds with $\gamma^B = \gamma^W=0$. Consequently, we have $$ \Phi := \int_0^T \upz{\hat{\pi}}_s\, ds =\Delta^{(0)},\quad \Delta^{(00)} :=\tfrac{1 }{1-p} {\mathbb E}^{\upz{\tilde{\PP}}}[\Lambda]. $$ To verify that \eqref{exp_reg} holds, we can use H\"older's inequality (twice) with exponents $-1/q$ and $(1-p)$ to see \begin{align*} {\mathbb E}^{\upz{\tilde{\PP}}}\left[e^{-\frac12 \varepsilon^2 \frac{p}{(1-p)^2}\Lambda + q\varepsilon \int_0^T \frac1{\sqrt{F_t}} dB^{\upz{\tilde{\PP}}}_t}\right]&\l {\mathbb E}^{\upz{\tilde{\PP}}}\left[e^{-\frac12 \varepsilon^2 (p+q)\Lambda}\right]^{\frac1{1-p}}\\ &\le \frac1{qv^{(0)}}{\mathbb E}\left[e^{-\frac12 \varepsilon^2 (1-p)(p+q)\Lambda}\right]^{\frac1{(1-p)^2}}, \end{align*} which is finite for $\varepsilon>0$ small enough. This allows Theorem \ref{thm:main3} to be invoked for $\varepsilon>0$ small enough. The second-order optimal controls $(\tilde{\pi},\tilde{\nu})$ are then well defined by \eqref{optimal_control_correction} and \eqref{optimal_dual_control_correction}, and read \begin{align} \tilde{\pi} := \hat{\pi}^{(0)} + \varepsilon \tfrac{\lambda'}{1-p},\quad \tilde{\nu} := \hat{\nu}^{(0)}. \end{align} Table 3 is the analogue of Table 2 for the extended affine model with parameters taken from Figure 4 in Section 3.3 in \cite{LarMun2012}. The methodology and the simulated quantities are the same as for Table 2. \newpage \begin{center} \begin{tabular}{cc||ccc} $\varepsilon$ & $F_0$ & $\text{CE}^{(\varepsilon)}(\hat{\pi}^{(0)})$ & LB & UB \\ \hline\hline $0.10$ & $0.01$ & $[1.724, 1.726]$& $[10.159, 10.399]$&$[10.226, 10.481]$ \\ $0.05$& $0.01$ & $[1.342, 1.343]$& $[2.141, 2.151]$&$[2.131, 2.149]$ \\ $0.01$& $0.01$& $[1.097, 1.098]$& $[1.118, 1.119]$&$[1.117, 1.120]$ \\ \hline $0.10$& $0.05$& $[1.728, 1.729]$& $[9.660, 9.877]$&$[9.766, 10.000]$ \\ $0.05$& $0.05$& $[1.344, 1.345]$& $[2.105, 2.115]$&$[2.102, 2.120 ]$ \\ $0.01$& $0.05$& $[1.099, 1.100]$& $[1.119, 1.121]$&$[1.117, 1.121]$ \\ \hline \end{tabular} \begin{quotation} \footnotesize \textbf{Table 3.} $95\%$-confidence intervals for certainty equivalents for the upper and lower bounds as well as the base model optimizer $\hat{\pi}^{(0)}$ for the extended affine model. The parameter values are $\kappa := 5$, $\theta := 0.0169$, $\beta := -0.1$, $\gamma :=0.1744$, $p:=-1$, and $T:=10$. The numbers are based on MC simulation using Euler's scheme with one million paths each with time-step size $0.001$. \end{quotation} \end{center} The zeroth order approximation $\text{CE}^{(0)}(\hat{\pi}^{(0)})$ produces the certainty equivalent values $$ \text{CE}^{(0)}(\hat{\pi}^{(0)}) = 1.043 \; (F_0=0.01),\quad\text{and }\quad \text{CE}^{(0)}(\hat{\pi}^{(0)}) = 1.045 \;(F_0=0.05). $$ Perhaps even more than in the Kim-Omberg model, the numbers in Table 3 above illustrate the superiority of the second-order approximations (columns 4 and 5) over its first-order version (column 3) as well as the zeroth order values reported above. Again, the bounds in Table 3 appear quite tight when compared to the first-order approximations for moderate values of $\varepsilon$. \bibliographystyle{amsalpha}
\section{Introduction} Nearly localized $f$ electrons in a metal interact with conduction electrons to cause a variety of interesting behaviors. By hybridization between $f$ and conduction electrons, each $f$ electron tends to form a spin-singlet state with conduction electrons, which is called the Kondo effect. If the crystalline-electric-field ground state makes a doublet with even number of $f$ electrons per site, as in Pr$^{3+}$- or U$^{4+}$-based systems, the two degenerate states is not connected by the time reversal, in contrast with a single spin. Such pair of states are called a non-Kramers doublet. It is known that the Kondo effect in the impurity system with a non-Kramers doublet causes peculiar phenomena such as a non-Fermi liquid ground state \cite{nozieres, cox, cox98}, where the entropy remains finite even at zero temperature. In contrast with a single impurity, the lattice of these non-Kramers doublets should undergo some ordering to remove the entropy of the whole system. There has been a long-standing interest in the resultant electronic order, especially in relation to exotic behaviors in non-Kramers systems such as UBe$_{13}$ \cite{ott, shimizu}, and more recently in PrIr$_{2}$Zn$_{20}$ and PrTi$_{2}$Al$_{20}$ \cite{onimaru, sakai, matsubayashi, tsujimoto}. The common features observed in these systems are the non-Fermi liquid behavior and superconductivity at low temperatures. To investigate the characteristic electronic order in non-Kramers doublet systems, we take the two-channel Kondo lattice as the simplest model \cite{emery93, jarrell97}. We have already demonstrated \cite{hoshino11,hoshino13,hoshino14} that the two-channel Kondo effect realizes exotic ground states with ``composite order'', where the order parameter is not described by one-body quantities such as magnetization or density, but by combination of $f$- and conduction-electron degrees of freedom. The diagonal composite order is identified as itinerant multipoles, and the off-diagonal one is a superconducting state with staggered order parameter. It has also been shown that a composite order is equivalent to ``odd-frequency order'' with a one-body but time-dependent order parameter of conduction electrons \cite{emery92, balatsky}. This concept is useful especially in detection of the instability toward composite orders. Furthermore, one of the authors \cite{hoshino14-2} has recently given a simple one-body picture of the orders by using a mean-field theory together with consideration of the SO(5) symmetry. In this paper we focus on collective excitation (Goldstone) modes that emerge from both diagonally and off-diagonally ordered states. These collective modes are gapless because the broken symmetry in either case is continuous; SU(2) in the diagonal order, and U(1) in the off-diagonal order. The detection of the peculiar Goldstone mode should provide clear evidence of the composite order. The collective modes couples either with magnetic field (diagonal order), or electric field (off-diagonal order). Hence powerful experimental probes such as neutrons, ultrasound or light can be used for the detection. In deriving the spectrum of the collective modes, we employ a mean-field theory and the random phase approximation. Before discussing the spectrum, we begin in the next section with reviewing the electronic orders and their symmetry in the two-channel Kondo lattice. In Sec. 3, we approach the ordered states from normal state at high temperatures by looking at the relevant response function. Although the response function probing the composite order is extremely complicated, the transition temperature can be derived by exploiting the equivalence to the odd-frequency one-body order. In Sec. 4 we proceed to the mean-field description of odd-frequency orders. The diagonally ordered state has a broken channel symmetry, and the corresponding mean-field is a hybridization between fictitious local fermions and conduction electrons in a particular channel. This hybridization results in a Kondo insulating state for one of the two channels, while the other channel remains a Fermi liquid. In the superconducting state, on the other hand, it is essential that we have two kinds of mean fields; hybridization in a particular channel, and pairing with fictitious $f$ electrons in the other channel. With these preliminaries, we derive in Sec. 5 the spectra of the Goldstone modes both in diagonally and off-diagonally ordered phases. The final section summarizes the paper, and discusses relevance of our results to actual materials. \section{Composite order parameters in Kondo lattice with non-Kramers doublets} Let us consider a non-Kramers $\Gamma_3$ doublet in cubic symmetry for the ground state of the $f^2$ configuration. We regard the doublet as pseudospin states described by the operator $\bm S$. We assume that they couple to conduction electrons with the $\Gamma_8$ representation, which have both pseudospin ($\sg$) and real spin ($\al$) degrees of freedom. This system is described by the two-channel Kondo lattice Hamiltonian given by \begin{align} {\cal H} = - t \sum_{\la ij \ra\al\sg} c^\dg_{i\al\sg } c_{j\al\sg} - \mu \sum_{i\al\sg} c^\dg_{i\al\sg } c_{i\al\sg} + J \sum_{i\al} \bm S_i \cdot \bm s_{ci\al} , \label{2chKL} \end{align} where $c_{i\al\sg}$ is the annihilation operator of conduction electron with channel $\al=1,2$ and pseudospin $\sg=\ua,\da$ at site $i$. The summation with respect to $\la ij \ra$ is taken over the nearest-neighbor sites on the bipartite lattice. We have also introduced pseudospin operator for conduction electrons as $\bm s_{ci\al} = \tfrac 12 \sum_{\sg\sg'} c^\dg_{i\al\sg} \bm \sg_{\sg\sg'} c_{i\al\sg'}$. Note that the channel $\al$ physically describes the real spin (or magnetic) degrees of freedom. The disordered state of this model cannnot be a Fermi liquid, since the localized spins are overscreened by conduction electrons due to the presence of two channels \cite{cox98}. This means that the disordered state has a residual entropy and cannot be the ground state. The simplest way to resolve the entropy is an ordering by the RKKY interaction among the localized pseudospins. In our previous studies, we have also found other peculiar ordered states with composite order parameters. We have both diagonal and off-diagonal composite orders, each of which is characterized by the following two-body quantities \cite{hoshino14}: \begin{align} \Psi^z &= \sum_{i\al\al'\sg\sg'} c^\dg_{i\al\sg} \sigma^z_{\al\al'}[\bm S_i \cdot \bm \sg_{\sg\sg'}] c_{i\al'\sg'} = 2 \sum_{i} \bm S_{i} \cdot (\bm s_{ci1} - \bm s_{ci2}) ,\label{Psi^z} \\ \Phi^+ &= \sum_{i\al\al'\sg\sg'} c^\dg_{i\al\sg} \epsilon_{\al\al'}[\bm S_i \cdot (\bm \sg \epsilon)_{\sg\sg'}] c^\dg_{i\al'\sg'} \, \epn^{\imu \bm Q\cdot \bm R_i} \end{align} where $\epsilon = \imu \sg^y$ is the antisymmetric unit tensor. Here $\Psi^z$ describes the channel symmetry breaking, and equivalent quantities $\Psi^x$ and $\Psi^y$ can be defined by changing the component of the Pauli matrix. On the other hand, $\Phi^+$ describes the channel-singlet/pseudospin-singlet pairing, and the equivalent quantity $\Phi^- \equiv (\Phi^+)^\dagger$ can be introduced. The phase factor $\exp(\imu \bm Q\cdot \bm R_i)$ becomes $\pm 1$ according to whether $\bm R_i$ belongs to A or B sublattice, and describes the staggered order. In the three-dimensional cubic lattice, for example, we take $\bm Q = (\pi, \pi, \pi)$ with lattice constant set to unity. These diagonal and off-diagonal orderings with five (3+2) independent components are degenerate at half filling. To demonstrate the degeneracy, let us introduce a unitary transformation $\mathscr{U}$ as follows: \begin{align} \mathscr{U} c_{i1\sg} \mathscr{U}^{-1} &= \frac 1 {\sqrt 2} \sum_{\sg'} (\delta_{\sg\sg'}c_{i1\sg} + \epsilon_{\sg\sg'} \epn^{\imu \bm Q \cdot \bm R_i} c^\dg_{i2\sg'} ) , \label{eq:unitary_transf1}\\ \mathscr{U} c_{i2\sg} \mathscr{U}^{-1} &= \frac 1 {\sqrt 2} \sum_{\sg'} (\delta_{\sg\sg'}c_{i1\sg} - \epsilon_{\sg\sg'} \epn^{\imu \bm Q \cdot \bm R_i} c^\dg_{i2\sg'} ) , \label{eq:unitary_transf2}\\ \mathscr{U} \bm S_{i} \mathscr{U}^{-1} &= \bm S_{i} , \label{eq:unitary_transf3} \end{align} which leaves the Hamiltonian invariant at half filling: $[\mathscr{U}, {\cal H}_{\mu=0}] = 0$. On the other hand, we obtain the relation \begin{align} \mathscr{U} \Psi^z \mathscr{U}^{-1} &= \Phi^x , \end{align} where $\Phi^x = (\Phi^+ + \Phi^-)/2$. Thus it is shown that the ordered states with $\Psi^z $ and $\Phi^x$ are degenerate. In a similar manner, we can demonstrate the degeneracy among other components by a rotation in the five-dimensional space of the order parameters, which is identical to the SO(5) representation or its isomorphic Sp(2) representation \cite{affleck}. For general filling with $\mu\neq 0$, we no longer have the particle-hole symmetry, and the diagonal and off-diagonal orders have different transition temperatures. It has been demonstrated numerically that the superconducting order is the most stable for a density range off half-filling \cite{hoshino14}. In the rest of the paper, we concentrate on the half-filled case with $\mu=0$, neglecting the trivial pseudospin order caused by the RKKY interaction. Owing to the SO(5) symmetry at $\mu=0$, information about the diagonal order gives equivalent knowledge about the off-diagonal order. \section{Odd-frequency susceptibility and instability toward composite order} In order to discuss the composite channel order, it is convenient to begin with the standard instability theory for the second-order phase transition in electron systems. We first introduce a homogeneous channel moment operator by \begin{align} m^z(\tau, \tau') = \sum_{i\al\al'\sg} c_{i\al\sg}^\dg (\tau) \sg^z_{\al\al'} c_{i\al'\sg} (\tau') , \end{align} where ${\cal O}(\tau) = \epn^{\tau {\cal H}} {\cal O} \epn^{-\tau {\cal H}}$ is the Heisenberg picture with imaginary time. We allow for $\tau\neq \tau'$ for later convenience. For the ordinary channel-ordered state, the order parameter is simply given by $\la m^z (0,0)\ra$. Let us assume that the channel order occurs by a second-order phase transition, which is signaled by divergence of the static susceptibility: \begin{align} \chi^{zz} = \int_0^\beta \la T_\tau m^z (\tau, \tau) m^z (0,0) \ra \, \diff \tau , \label{eq:chan_suscep} \end{align} where $T_\tau$ is the imaginary-time ordering operator. The susceptibility can be rewritten in terms of the two-particle Green function. Namely we define \begin{align} \chi^{zz} (\tau_1, \tau_2, \tau_3, \tau_4) &= \la T_\tau m^z (\tau_1, \tau_2) m^z (\tau_3, \tau_4) \ra , \label{eq:tpgf} \end{align} and its Fourier transform by \begin{align} \chi^{zz} (\imu\ep_n, \imu\ep_{n'}) = \frac{1}{\beta^2} \int_0^\beta \diff\tau_1\cdots \diff\tau_4 \ \chi^{zz} (\tau_1, \tau_2, \tau_3, \tau_4) \, \epn^{\imu\ep_n(\tau_2 - \tau_1)} \epn^{\imu\ep_{n'}(\tau_4 - \tau_3)} , \end{align} where $\ep_n = (2n+1)\pi/\beta$ is the fermionic Matsubara frequency. In terms of of this quantity, Eq.~\eqref{eq:chan_suscep} can be written as \begin{align} \chi^{zz} = \frac{1}{\beta} \sum_{nn'} \chi^{zz} (\imu\ep_n, \imu\ep_{n'}) . \end{align} Let us now turn to the composite ordering where the order-parameter is a two-body quantity. This means that the corresponding susceptibility is made of four-body quantity, which is extremely cumbersome to calculate, if not impossible. Instead, we notice the relation \begin{align} \left. \frac{\partial m^z(\tau, \tau')}{\partial \tau'} \right|_{\tau,\tau'=0} \equiv \tilde m^z (\tau, \tau') = \frac{J}{2} \Psi^z - t \sum_{\la ij\ra\al\al'\sg} c_{i\al\sg}^\dg \sg^z_{\al\al'} c_{j\al'\sg} , \label{eq:deriv_chan_mom} \end{align} where the right-hand side (RHS) includes $\Psi^z$. Hence the left-hand side of Eq.~\eqref{eq:deriv_chan_mom} describes the composite order parameter. Namely, the instability toward the composite channel order can be detected from the following susceptibility: \begin{align} &\tilde \chi^{zz} = \int_0^\beta \la T_\tau \tilde m^z (\tau, \tau) \tilde m^z (0,0) \ra \, \diff \tau = - \frac{1}{\beta} \sum_{nn'} g(\ep_n) g(\ep_{n'}) \chi^{zz} (\imu\ep_n, \imu\ep_{n'}) + 2\la {\cal H} \ra , \label{eq:tilde_suscep} \end{align} where $g(\ep_n) = \ep_n \epn^{\imu \ep_n \eta}$ with $\eta = +0$. The form factor $g(\ep_n)$ has appeared because of the time-derivative, and the factor $\epn^{\imu \ep_n \eta}$ is necessary to attain the convergence. The term $\la {\cal H} \ra$ in the RHS originates from the time-ordering operator in the two-particle Green function \eqref{eq:tpgf}, which can be neglected as long as we are concerned about the divergence of $\tilde \chi^{zz}$. For practical calculation, the form factor $g(\ep_n)$ as it stands is awkward since the high-frequency part in the summation has a large contribution, even though it remains finite at any temperature due to the presence of $\epn^{\imu \ep_n \eta}$. Since we are concerned only with divergence of the Eq.~\eqref{eq:tilde_suscep}, we replace $g(\ep_n)$ by \begin{align} g(\ep_n) &\longrightarrow {\rm sgn\,} \ep_n , \end{align} with which we no longer need the convergence factor. In the imaginary-time domain, it is equivalent to the following replacement: \begin{align} \tilde \chi^{zz} \longrightarrow - \frac{1}{\beta^2} \int_0^\beta \diff \tau_1 \cdots \tau_4 \, s(\tau_1 - \tau_2) s(\tau_3-\tau_4) \la T_\tau m^z (\tau_1, \tau_2) m^z (\tau_3, \tau_4) \ra , \end{align} where \begin{align} s(\tau) = \imu \, {\rm cosec} ( {\pi \tau}/\beta). \end{align} With this replacement, the response function becomes quantitatively different from the original susceptibility. Nevertheless, the divergence occurs at the same transition temperature because the replacement causes only a finite difference. Since the form factor relevant to the composite order is an odd function in frequency, $\tilde \chi^{zz}$ is called an odd-frequency susceptibility. To make contrast, ordinary susceptibility such as Eq.~\eqref{eq:chan_suscep} may be called even-frequency susceptibility. Thus ordinary and composite channel-symmetry breakings are detected by even- and odd-frequency susceptibilities, respectively, both of which can be calculated from the two-particle Green function. We can apply the same procedure to detect the composite pairing state. Among various internal symmetries of the pairing, the most stable is the channel-singlet/pseudospin-singlet pairing \cite{hoshino14}. Hence, the relevant quantity is now \begin{align} p^+ (\tau, \tau') = \sum_{i\al\al'\sg\sg'} c^\dg_{i\al\sg}(\tau) \epsilon_{\al\al'} \epsilon_{\sg\sg'} c^\dg_{i\al'\sg'} (\tau') \, \epn^{\imu \bm Q\cdot \bm R_i} \end{align} instead of $m^z$. We have already evaluated numerically both diagonal and off-diagonal odd-frequency susceptibilities by combining the continuous-time quantum Monte Carlo (CT-QMC) and the dynamical mean-field theory (DMFT)\cite{hoshino11, hoshino14}. We have indeed found the divergence corresponding to the onset of diagonal and off-diagonal composite orders. \section{Effective Hamiltonian with pseudofermions} From numerical results \cite{hoshino11, hoshino14}, we have realized that the low-energy states in the ordered phase can be described in terms of fermionic excitations that originate not only from conduction bands but also from localized states forming the non-Kramers doublet. Analogous situation has long been recognized in the standard Kondo lattice, where the local spins acquire the fermionic character by the Kondo effect. The paramagnetic fixed point of the Kondo lattice can be described by hybridized bands with strongly renormalized parameters, which comes from Kondo resonance at each site. Let us first construct the effective Hamiltonian for the diagonal composite order described by $\Psi^z$ of Eq.~\eqref{Psi^z}. In the ordered phase, the conduction electrons in channel $\al=1$ forms a Kondo insulator with hybridized band, while the channel $\al=2$ remains a Fermi liquid decoupled form localized pseudospins. This difference between the two channels is reflected in the order parameter $\la \Psi^z\ra = 2\sum_{i} \la \bm S_i \cdot (\bm s_{ci1} - \bm s_{ci2}) \ra$. In analogy with the ordinary Kondo insulator \cite{riseborough, zhang}, we first rewrite the localized pseudospin operator as \begin{align} &\bm S_i = \frac 1 2 \sum_{i\sg\sg'} f^\dg_{i\sg} \bm \sg_{\sg\sg'} f_{i\sg'} , \\ & \sum_{\sg} f^\dg_{i\sg} f_{i\sg} = 1 . \label{eq:local_constraint} \end{align} Note that fermionic operator $f_{i\sg}$ does not come from original $f$ electrons forming the non-Kramers doublet. Hence we call these pseudofermions. In the mean-field theory, the local operator constraint \eqref{eq:local_constraint} is replaced by a much looser constraint for the average of the operator. In the composite order with $\Psi^z$, we replace the interaction term with channel $\alpha=1$ as \begin{align} J \bm S_i \cdot \bm s_{ci1} \longrightarrow V_1 \sum_\sg f^\dg_{i\sg} c_{i1\sg} + {\rm h.c. } \end{align} where the mean field \begin{align} V_1= - J \la c^\dg_{i1\sg} f_{i\sg} \ra \label{V_1} \end{align} is independent of the site. On the other hand, there is no mean field from the interaction $J \bm S_i \cdot \bm s_{ci2}$ because the channel 2 does not form the hybridized band. Note that we would obtain $3J/4$ instead of $J$ in Eq.~\eqref{V_1} in the faithful decoupling procedure. However we prefer the present choice since it gives the correct Kondo scale \cite{lacroix-cyrot1983}. In this way we obtain the mean-field Hamiltonian \begin{align} {\cal H}_{\Psi^z} &= \sum_{\bm k\al\sg} \ep_{\bm k}c^\dg_{\bm k\al\sg} c_{\bm k\al\sg} + V_1\sum_{\bm k\al\sg } \left( f^\dg_{\bm k\sg} c_{\bm k1\sg} + c^\dg_{\bm k1\sg} f_{\bm k\sg} \right) , \label{eq:MF_ham} \end{align} where the operators in the momentum space are defined by the Fourier transform of the real-space operators. The single-particle spectra consist of three branches, which are given by \begin{align} E_{\bm k\pm} &= \frac 1 2 \left[ \ep_{\bm k} \pm \sqrt{\ep_{\bm k}^2 + 4V_1^2} \right] , \label{eq:sp_energy1} \\ E_{\bm k0} &= \ep_{\bm k} , \label{eq:sp_energy2} \end{align} where $E_{\bm k0}$ describes quasi-particles in channel $\al=2$. These branches qualitatively reproduce the results obtained by the DMFT at low energies. If we solve the mean-field equation \eqref{V_1} self-consistently, the characteristic energy scale $\Gamma$ such as the transition temperature and ground-state energy is obtained as $\rho (0) \Gamma \sim \epn^{- 1/\rho(0) J}$ where $\rho(0)$ is the density of states at the Fermi level. This gives usual expression for the Kondo temperature. We note that the approximation here is valid only at sufficiently low temperatures, since the disordered state of the two-channel Kondo lattice is a non-Fermi liquid which cannot be described by the present theory. Next we turn to the effective Hamiltonian in the ordered phase with $\Phi^x$. Because of the SO(5) symmetry, which applies not only to the ground state but all excited states, the total single-particle spectrum \begin{align} \rho (\bm k,\ep) = -\frac 1{2\pi} \imag\, {\rm Tr}\, \hat{\bf G}(\bm k, \ep+\imu\eta) \end{align} is the same in ordered states either with $\Psi^z$ or $\Phi^x$. Here the Green function $\hat{\bf G}(\bm k, \ep+\imu\eta)$ is an $8\times8$ matrix in the generalized Nambu space with components; $(c^\dagger _{\al\sg}, c_{\al'\sg'})$ with $\al,\al' =1,2$ and $\sg,\sg' =\pm 1$. To be more specific, using the unitary transformations defined in Eqs.~(\ref{eq:unitary_transf1}--\ref{eq:unitary_transf3}), we can derive the effective Hamiltonian for composite pairing state. Let us first regard $V_1$ as independent of $\mathscr{U}$. Then the result of transformation is given by \begin{align} \mathscr{U} {\cal H}_{\Psi^z} \mathscr{U}^{-1} \equiv {\cal H}_{\Phi^x} = \sum_{\bm k\al\sg} \ep_{\bm k}c^\dg_{\bm k\al\sg} c_{\bm k\al\sg} + \frac{V_1}{\sqrt 2}\sum_{\bm k\sg\sg' } \left( \delta_{\sg\sg'} f^\dg_{\bm k\sg} c_{\bm k1\sg} + \epsilon_{\sg\sg'} f^\dg_{\bm k\sg} c^\dg_{-\bm k-\bm Q, 2\sg'} + {\rm h.c.} \right). \end{align} If we start from the original model given by Eq.~\eqref{2chKL}, we may decouple the interaction part so as to allow for a new nonzero average such as $\la c_{i 2\sg} f_{i\sg} \ra \neq 0$. By setting equal magnitudes to diagonal and off-diagonal mean fields, we obtain the result identical to ${\cal H}_{\Phi^x}$. As can be checked by direct calculation, the single-particle spectra of ${\cal H}_{\Phi^x}$ are the same as Eqs.~\eqref{eq:sp_energy1} and \eqref{eq:sp_energy2}. However, the wave function of each excitation is of course different. As a result, ${\cal H}_{\Psi^z} $ and ${\cal H}_{\Phi^x}$ have the same specific heat, but show very different electromagnetic properties. \section{Collective excitation modes} Now we discuss collective excitations in the composite channel ordered state with $\Psi^z$. Within the equation of motion method, we include self-consistently induced field that corresponds to vertex corrections in the Bethe-Salpeter equation \cite{anderson}. In order to derive collective excitations from the ordered state, it is convenient to add an (infinitesimal) channel flipping field $h_c$ with wave number $\bm q$: \begin{align} {\cal H}_{\rm ext}(\bm q) &= - h_c \sum_{\bm k\sg } \left( c^\dg_{\bm k+\bm q,1 \sg} c_{\bm k2\sg} + c^\dg_{\bm k2\sg} c_{\bm k+\bm q,1 \sg} \right) . \end{align} The channel field in the non-Kramers doublet corresponds to a modulating magnetic field in the $xy$-plane. With this external field, hybridization in channel $\al=2$ is induced infinitesimally. In the random phase approximation, the induced hybridization is accounted for by the Hamiltonian \begin{align} {\cal H}_{\rm ind}(\bm q) &= V_2 \sum_{\bm k\sg } \left( f^\dg_{\bm k+\bm q, \sg} c_{\bm k2\sg} + c^\dg_{\bm k2\sg} f_{\bm k+\bm q, \sg} \right), \\ V_2 = &-\frac{J}{N} \sum_{\bm k} \la c^\dg_{\bm k2\sg} f_{\bm k+\bm q,\sg} \ra , \end{align} where $N = \sum_i 1$ is the total number of sites, and $V_2 \sim O(h_c)$. With ${\cal H} = {\cal H}_{\Psi^z} + {\cal H}_{\rm ext} + {\cal H}_{\rm ind}$, the Heisenberg equations of motion $\imu\partial_t {\cal O} = [{\cal O}, {\cal H}]$ are explicitly written as \begin{align} \imu \partial_t (c^\dg_{\bm k+\bm q, 1\sg} c_{\bm k2\sg}) &= (- \ep_{\bm k + \bm q}+\ep_{\bm k}) c^\dg_{\bm k+\bm q, 1\sg} c_{\bm k2\sg} - V_1 f^\dg_{\bm k+\bm q, \sg} c_{\bm k2\sg} + V_2 c^\dg_{\bm k+\bm q, 1\sg} f_{\bm k+\bm q, \sg} \nonumber \\ &\ \ + h_c c^\dg_{\bm k 2\sg} c_{\bm k2\sg} - h_c c^\dg_{\bm k+\bm q, 1\sg} c_{\bm k+\bm q, 1\sg} , \\ \imu \partial_t (f^\dg_{\bm k+\bm q, \sg} c_{\bm k2\sg}) &= \ep_{\bm k} f^\dg_{\bm k+\bm q, \sg} c_{\bm k2\sg} - V_1 c^\dg_{\bm k+\bm q, 1\sg} c_{\bm k2\sg} - V_2 c^\dg_{\bm k 2\sg} c_{\bm k2\sg} + V_2 f^\dg_{\bm k+\bm q, \sg} f_{\bm k+\bm q, \sg} \nonumber \\ &\ \ - h_c f^\dg_{\bm k+\bm q, \sg} c_{\bm k+\bm q, 1\sg} . \end{align} If the external field is oscillating with frequency $\omega$, the Fourier components are determined by \begin{align} &\begin{pmatrix} \la c^\dg_{1\sg} c_{2\sg} \ra (\bm q, \omega) \\ \la f^\dg_{\sg} c_{2\sg} \ra (\bm q, \omega) \end{pmatrix} = \begin{pmatrix} \chi^{11}_{\bm q}(\omega) & \chi^{12}_{\bm q}(\omega)\\ \chi^{21}_{\bm q}(\omega) & \chi^{22}_{\bm q}(\omega)\\ \end{pmatrix} \begin{pmatrix} h_c \\ -V_2 \end{pmatrix}, \\ &\begin{pmatrix} \chi^{11} & \chi^{12}\\ \chi^{21} & \chi^{22}\\ \end{pmatrix} = - \frac 1 N \sum_{\bm k} \begin{pmatrix} \omega+\ep_{\bm k+\bm q}-\ep_{\bm k} & V_1\\ V_1 & \omega - \ep_{\bm k} \\ \end{pmatrix}^{-1} \begin{pmatrix} n_{c1, \bm k+\bm q} - n_{c2\bm k} & X_{\bm k+\bm q}\\ X_{\bm k+\bm q} & n_{f, \bm k+\bm q} - n_{c2\bm k} \end{pmatrix}, \label{eq:def_chi} \end{align} where the quantities in the RHS are evaluated from the Hamiltonian ${\cal H}_{\Psi^z}$ as \begin{align} X_{\bm k} &= \la f^\dg_{\bm k\sg} c_{\bm k1\sg} \ra = V_1 [f(E_{\bm k+}) - f(E_{\bm k-})] / (E_{\bm k+}-E_{\bm k-}) , \\ n_{f\bm k} &= \la f^\dg_{\bm k\sg} f_{\bm k\sg} \ra = [- E_{\bm k-}f(E_{\bm k+}) + E_{\bm k+} f(E_{\bm k-})] / (E_{\bm k+}-E_{\bm k-}) , \\ n_{c1\bm k} &= \la c^\dg_{\bm k1\sg} c_{\bm k1\sg} \ra = [E_{\bm k+}f(E_{\bm k+}) - E_{\bm k-}f(E_{\bm k-})] / (E_{\bm k+}-E_{\bm k-}) , \\ n_{c2\bm k} &= \la c^\dg_{\bm k2\sg} c_{\bm k2\sg} \ra = f(E_{\bm k0}) , \end{align} with the Fermi distribution function $f(\ep) = 1 / (\epn^{\beta\ep}+1)$. Using these quantities, the relevant dynamical susceptibility is derived as \begin{align} \chi_{\bm q}(\omega) &= \la c^\dg_{1\sg} c_{2\sg} \ra (\bm q, \omega)/h_c = \frac{\chi^{11} - J (\chi^{11}\chi^{22} - \chi^{12}\chi^{21})} {1 - J \chi^{22}} , \label{eq:dynamical_sus1} \end{align} whose pole determines the dispersion relation of the Goldstone modes. Assuming small $|{\bm q}|$ and $\omega$, we expand the denominator as \begin{align} 1-J\chi^{22}_{\bm q} (\omega) &= c_0 (\bm q) + c_2 (\bm q) \omega^2 + O(\omega^4) . \label{eq:expand_pole} \end{align} The odd-order terms of $\bm q$ and $\omega$ vanish because of the spatial inversion and particle-hole symmetries, respectively. In the lowest-order with respect to $\bm q$ and $\omega$, each coefficient is evaluated at $T=0$ as \begin{align} c_0(\bm q) &= - \frac{J}{2NV_1^4} \sum_{\bm k} (\bm v_{\bm k} \cdot \bm q)^2\left[ \frac{\ep_{\bm k}^4 + 6V_1^2\ep_{\bm k}^2 + 6V_1^4}{(\ep_{\bm k}^2 + 4V_1^2)^{3/2}} - |\ep_{\bm k}| - V_1^2 \delta (\ep_{\bm k}) \right] + O(\bm q^4), \label{eq:c0} \\ c_2(\bm q) &= \frac{J}{2NV_1^4} \sum_{\bm k} \left[ |\ep_{\bm k}| - \frac{\ep_{\bm k}^2 + 2V_1^2}{\sqrt{\ep_{\bm k}^2 + 4V_1^2}} \right] +O(\bm q^2). \end{align} Here we have defined the velocity by $\bm v_{\bm k} = \partial \ep_{\bm k} / \partial \bm k$. For sufficiently small $V_1$, the integrands are sharply peaked at the Fermi level, and hence we can represent the density of states by its values near the Fermi level. Namely, we use the following approximations \begin{align} \tilde \rho (\ep) &= \frac 1 N \sum_{\bm k} \bm v_{\bm k}^2 \delta (\ep - \ep_{\bm k}) = \tilde \rho (0) + \frac 1 2 \tilde \rho''(0) \ep^2 + O(\ep^4) ,\label{eq:v-DOS} \\ \rho (\ep) &= \frac 1 N \sum_{\bm k} \delta (\ep - \ep_{\bm k}) = \rho (0) + O(\ep^2) .\label{eq:DOS} \end{align} In Eq.~\eqref{eq:v-DOS}, we need terms up to $O(\ep^2)$ since contribution of $\tilde \rho (0)$ is cancelled in Eq.~\eqref{eq:c0}. Thus the energy integration can be performed. We note that $\tilde \rho''(0)$ is estimated as $\tilde \rho''(0) \sim - v_{\rm F}^2 \rho(0) / D^2$ with $v_{\rm F}^2 = \la \bm v_{\bm k}^2 \ra_{\rm FS}$ being averaged on the Fermi surface and $D$ being the energy with the order of the band width. The dispersion relation $\omega = \omega_{\bm q}$ is determined by putting Eq.~\eqref{eq:expand_pole} equal to zero, and we obtain \begin{align} \omega_{\bm q} = \frac{v_{\rm F}}{\sqrt{d}} \left( \frac{V_1}{2D} \right) \, |\bm q| , \label{eq:gs_dispersion} \end{align} where $d=3$ is the dimension of the system. The velocity of the collective modes is proportional to the Fermi velocity and the magnitude of the mean-field. The factor $V_1/D$ arises due to hybridization between the conduction electrons and localized pseudofermions, which modifies the Fermi velocity $v_{\rm F}$ of conduction electrons. The combination of the Fermi velocity and the order parameter in giving the velocity is a characteristic feature in the present ordered state. Namely, in itinerant magnetism or superfluid in the weak-coupling limit, the velocity of the Goldstone mode is close to the Fermi velocity. On the other hand, the velocity in localized antiferromagnet is proportional to the exchange interaction times the spontaneous magnetization within the mean-field theory. \begin{figure}[t] \begin{center} \includegraphics[width=60mm]{gs_mode3.eps} \caption{(Color online) Two-particle spectrum probed by $\imag \chi_{\bm q = (q_x, 0, 0)} (\omega+\imu\eta)/\omega$. The parameters are chosen as $V_1 = 0.5$, $T=0.025$ and $\eta = 0.00375$ with $t=1$ being the unit of energy. } \label{fig:gs_mode} \end{center} \end{figure} The two-particle spectrum appears as peaked features in $\imag \chi_{\bm q}(\omega+\imu\eta)/\omega$ if scanned over the $(\bm q,\omega)$ plane. We assume the simple cubic lattice in three dimension, where the conduction band is described by \begin{align} \ep_{\bm k} = -2t (\cos k_x + \cos k_y + \cos k_z) \end{align} with the lattice constant being unity. Figure \ref{fig:gs_mode} shows the two-particle spectra calculated from Eq.~\eqref{eq:dynamical_sus1}. We can clearly see the low-lying collective mode located around $\bm q=\bm 0$. Experimentally, the channel moment corresponds to the spin magnetic moment of conduction electrons. Hence, if inelastic neutron scattering experiment is performed in a system with the composite channel order, the magnetic spectrum should look like Fig.~\ref{fig:gs_mode}. On the other hand, for composite pairing with the staggered ordering vector $\bm Q$, the spectrum of the Goldstone mode is derived simply by shifting the momentum as $\bm q \rightarrow \bm q + \bm Q$, provided the system is at half filling. The resultant dispersion relation $\omega = \omega'_{\bm q}$ is given by \begin{align} \omega'_{\bm q} = \omega_{\bm q - \bm Q} \end{align} where the RHS is given by Eq.~\eqref{eq:gs_dispersion} for small $|\bm q-\bm Q|$. Namely, the intensity map in the $(\bm q,\omega)$ space should again look like Fig.~\ref{fig:gs_mode} but the branch arises at $\bm q = \bm Q$. The collective mode in the composite pairing state can be detectable in the charge part of the dynamical susceptibilities. It is well known for ordinary superconductors that the charge Goldstone mode is absorbed into a plasmon by the long-ranged Coulomb interaction \cite{anderson}, which has singularity in the limit $\bm q\rightarrow \bm 0$. In the present staggered pairing state, on the other hand, the Coulomb interaction has no singularity at $\bm q = \bm Q$. Hence it is expected that the coupling of the collective excitation with the plasmon mode is weaker, or negligible. At low temperature, the inelastic light scattering experiment is one of candidates to detect this type of off-diagonal order. \section{Summary and Discussion} We have demonstrated in this paper that onset of the composite order parameter can be detected from divergence of the odd-frequency susceptibility, which is derived within the standard framework of two-particle Green functions. We have also shown that a mean-field theory with pseudofermions provides a much simpler description of composite orders in the two-channel Kondo lattice. The simplicity of the framework has made it possible to derive the spectrum of the collective excitations from the ordered state. In actual non-Kramers doublet systems, the particle-hole symmetry in the conduction bands can hardly be realized. In this case the diagonal composite order accompanies ferromagnetic moment of conduction electrons \cite{hoshino11}. The magnitude of the moment depends on degree of the broken particle-hole symmetry, but is expected to be tiny in general. Note that the present collective mode in the presence of spontaneous moment is different from the ordinary ferromagnetic magnon mode. Namely, the spectrum $\omega_{\bm q} \propto |\bm q|$ is in contrast with the magnon spectrum $\omega_{\bm q} \propto \bm q^2$. The latter is common to localized and itinerant ferromagnets with spin SU(2) symmetry. If there is a magnetic anisotropy caused, for example, by spin-orbit interaction in conduction bands, the collective mode in the composite diagonal order should acquire a gap in the spectrum. On the other hand, it has been demonstrated \cite{hoshino13,hoshino14} that the composite superconductivity is more stable than the diagonal order in the density range without the particle-hole symmetry. Hence we expect that actual competition in real materials should be between the ordinary quadrupole order by the RKKY interaction, and the composite superconductivity. If the composite superconductivity is indeed realized in non-Kramers doublet systems, the Goldstone mode should remain gapless even without the particle-hole symmetry. Furthermore, the gapless point corresponds to the boundary of the Brillouin zone. In the smaller Brillouin zone associated with the staggered order, the zone boundary is folded to the center of the Brillouin zone. It remains to see how the zone-folding effect affects the physical property of the collective mode. We expect that this collective mode may be probed by inelastic light scattering and other measurement that detects excitations with charge degrees of freedom. The observation clearly identifies the composite superconductivity, as distinct from the ordinary one. \section*{References}
\section{Introduction} \setcounter{equation}{0} \ Nonlinear optics has lead the way of several fundamental discoveries, elegant and fascinating one of these is the optical solitons. These solitons are extensively studied since last three decades not only because of their mathematical elegance but also due to applications in photonics and optical communications~\citep{1,2,3,4,5,6,7,8,9,10,11,12,13,14,15,16,17,18,19,20,21,22,23,24}. Optical beams and pulses spread during propagation due to self-diffraction or dispersion. This broadening can be arrested in optical nonlinear media resulting in stable optical beams or pulses whose spreading is exactly balanced by optical nonlinearity and these self trapped beams or pulses are known as optical solitons~\citep{2,3,5,7}. Optical spatial solitons are self-trapped optical beams whose spreading due to diffraction is exactly balanced by optical nonlinearity induced self lensing mechanisms.The optical beam induces a refractive index change in the medium thereby creating an optical waveguide which subsequently guides the beam. Thus, the beam first creates an waveguide and then self-trapped and guided in the waveguide created by itself. An optical temporal soliton on the otherhand is a nonspreading optical pulse whose broadening due to group velocity dispersion is exactly balanced by the nonlinearity induced self phase modulation. More explicitly, the group velocity dispersion produces a chirp in the propagating pulse which is balanced by the opposite chirp created by nonlinearity. Exact balance of these opposite chirps results in the formation of optical temporal solitons. \subsection{Optical Solitons or Optical Solitary Waves? } Way back in 1995, Zabusky and Kruskal~\citep{13} introduced the term solitons to reflect the particle like properties of stable self-trapped waves in nonlinear media. The motivation behind this was the fact that the shape of these self-trapped wavepackets remain intact even after collision. Historically, this term was reserved for those wavepackets which obey integrable partial differential equations that can be solved by inverse scattering theory~\citep{10,11,13}. Solitons, which are solutions of these equations, remain invariant even after collision since they undergo elastic collision. Optical beam and pulse propagation in Kerr nonlinear media are governed by partial differential equations which are integrable via inverse scattering technique, hence, optical solitons can be created in such media. However, most physical optical systems, for example photorefractive media, possess non-Kerr types of nonlinearity in which optical beams or pulses are described by dynamical equations that are though partial differential equations but not integrable by inverse scattering technique. Self-trapped solutions of these non-integrable equations are known as 'solitary waves'. These solitary waves are not solitons since collision between these solitary waves is not always elastic~\citep{8,9,14}. For example, under specific conditions, two photorefractive optical spatial solitary waves may coalesce to form a single solitary wave and a single solitary wave might undergo fission to give birth to new solitons. However, these solitary waves in many cases play extremely important role in several applications. In literature, despite their shortcoming, it is now quite common to loosely refer all self-trapped optical solitary waves as solitons regardless of whether they obey integrable partial differential equation or not. \subsection{Emergence of Photorefractive Optical Spatial Solitons } Guiding of optical beams in the self created waveguide was first pointed out by Askaryan in 1962~\citep{12}. Kelly~\citep{15} pointed out that in Kerr nonlinear media spatial solitons with two transverse dimensions(2D) are unstable and undergo catastrophic collapse. Soon it was realised that spatial solitons with only one transverse dimensions(1D) are stable in Kerr nonlinear media~\citep{15}. Subsequently, it was also realized that catastrophic collapse of 2D solitons could be avoided in an optical media possessing saturating nonlinearity~\citep{16}. Therefore, the idea of saturating nonlinearity was the key to the discovery of a plethora of optical spatial solitons.\ Inspite of the prediction of formation of stable 2D solitons in saturating nonlinear media, no progress was made in creating these solitons due to lack of identification of appropriate optical materials that possess saturating nonlinearity. The breakthrough was made by Segev et al~\citep{17,18} in 1990 with the identification of non-Kerr type of nonlinearity in photorefractive media and observation of optical spatial solitons due to such nonlinearities. Unlike Kerr solitons, the dynamics of photorefractive solitons, under a variety of photorefractive nonlinear effects, is governed by modified nonlinear Schr$\ddot{o}$dinger equations. Photorefractive nonlinearity is non-instantaneous, typically nonlocal and inherently saturable and spatial solitons in these media can be created at microwatt power level using very simple set up in the laboratory, while their formation time ranges from microseconds to minutes. \\\\ Photorefractive optical spatial solitons possess several unique properties, for example unlike Kerr solitons which undergo elastic collision, collisions of these solitons are inelastic, more diverse and interesting. The inelastic collision between two photorefractive spatial solitons may lead to soliton fusion or fission, with particle-like annihilation or birth of new solitons. Another unique feature is that one can create an induced waveguide using a weak soliton beam which subsequently can be used to guide another powerful beam at different wavelength at which the material is less photosensitive ~\citep{24,25}. It has been demonstrated that photorefractive spatial soliton-induced waveguides can be used for device applications such as directional couplers and high efficiency frequency converters ~\citep{26,27}. These unique features of photorefractive nonlinearity have opened up immense possibilities for new and novel devices, such as all optical switching and routing, steering, interconnects, parallel computing, optical storage etc~\citep{28,29,30,31} . They are also promising for experimental verification of theoretical models, since, they can be created at very low power. In the present article, we confine our discussion on the properties of photorefractive optical spatial solitons. This review is only partial description of the development of photorefractive optical spatial solitons, interested readers are referred to several excellent recent reviews~\citep{14,32,33,34,35}. \section{Photorefractive Materials } Exposure of a photorefractive (PR) material with optical field of nonuniform intensity leads to the excitation of charge carriers of inhomogeneous density. These charge carriers then migrate due to drift or diffusion or both and creates space charge field which subsequently modifies the refractive index of the material via either linear electro-optic or quadratic electro-optic effect ~\citep{30,31,35}. Change in the refractive index of certain electro-optic materials due to optically induced redistribution of charge carriers is known as photorefractive(PR) effect. Photorefractive effect has potential applications in holography~\citep{34}, optical phase conjugation~\citep{30,31,36}, optical signal processing and optical storage~\citep{31,34,35}. Generally, PR materials are classified in three different categories. Most commonly used PR materials are inorganic ferroelectrics, such as $LiNbO_{3}$, $KNbO_{3}$, $BaTiO_{3}$ etc. Recently, large number of experiments on spatial solitons have been performed~\citep{37} in centrosymmetric paraelectric potassium lithium tantalate niobate $(KLTN)$. Some selected semiconductors such as InP, GaAs, CdTe etc., also show PR property with large carrier mobility that produces fast dielectric response time, which is important for fast image processing. They have potential use in fast holographic processing of optical information. The third category PR materials are polymers which shows strong PR effect~\citep{34,35} at high applied voltage. PR pattern can be erased easily in polymers by decreasing applied voltage. \section{Protorefractive Optical Nonlinearity} The band transport model of Kukhtarev-Vinetskii~\citep{38} has been widely used to describe the theoretical foundation of PR nonlinearity. In view of this model, we assume a PR medium with completely full valance band and an empty conduction band which is illuminated by an optical field of nonuniforn intensity. It has both donor and acceptor centers, uniformly distributed, whose energy states lies somewhere in the middle of the band gap. The donor electron states are at higher energy in comparison to that of acceptor energy states. The nonuniform optical field excites unionized donors and creates charge carriers. These charge carriers move to the conduction band where they are free to move, to diffuse or to drift under the combined influence of self generated and external electric field and are finally trapped by the acceptors. During this process, some of the electrons are captured by ionized donors and thus are neutralized, and finally a steady state is reached with the creation of internal space charge field, that can be evaluated using donor ionization rate equation, electron continuity equation, current density $(J)$ equation, Poisson's equation and the charge density($\rho$) equation~\citep{14,17,18,19, 20,21,22,23,24,39,40} which are as follows: \begin{equation}\label{e1} \frac{\partial N_{D}^{+}}{\partial t}=(s_{i}I+\beta_{T})(N_{D}-N_{D}^{+})-\gamma _{R}N N_{D}^{+}, \ \end{equation} \begin{equation}\label{e2} \frac{\partial N}{\partial t}-\frac{\partial N_{D}^{+}}{\partial t} =\frac{1}{e}\overrightarrow{\nabla }.\overrightarrow{J}, \end{equation} \begin{equation}\label{e3} \overrightarrow{J}=e N\mu \overrightarrow{E} +k_{B}T\mu \overrightarrow{\nabla }N + k_{p}s_{i}\left(N_{D}-N_{D}^{+}\right)I\vec{c}, \end{equation} \begin{equation}\label{e4} \overrightarrow{\nabla}\cdot\epsilon \overrightarrow{E}=\rho , \end{equation} \begin{equation}\label{e5} \rho \left(\overrightarrow{r}\right)=e\left(N_{D}^{+}-N_{A}-N\right), \end{equation} where $N $, $N_{D}$, $N_{A}$ and $N_{D}^{+}$ are densities of electron, donor, acceptor, and ionized donor, respectively; $s_{i}$ is the photoexcitation cross section, $I$ is the intensity of light in terms of Poynting flux, $\beta_{T}$ and $\gamma _{R}$ are rate of thermal generation and electron trap recombination coefficient, respectively. $T$, $e$ and $\mu$ are respectively electron's temperature, charge and mobility; $k_{B}$ is the Boltzmann constant, $k_{p}$ is the photovoltaic constant and $\vec{c}$ is the unit vector in the direction of c-axis of the PR crystal. The current density arises due to the drift, diffusion and photovoltaic effect; $E$ is the sum of externally applied field and the generated space charge field $E_{sc}$. Most experimental investigations on PR solitons have been performed using one dimensional waves, therefore, it is appropriate to find material response in one transverse dimension (say x only). In the steady state equations (1)-(4) reduces to: \begin{equation}\label{e6} s_{i}(I+I_{d})(N_{D}-N_{D}^{+})-\gamma_{R}NN_{D}^+=0, \end{equation} \begin{equation}\label{e7} \frac{\partial E_{sc}}{\partial x}=\frac{e}{\epsilon_{0}\epsilon_{r}}(N_{D}^{+}-N_{A}-N), \end{equation} \begin{equation}\label{e8} J=e N\mu E_{sc} +k_{B}T\mu \frac{\partial N}{\partial x} + k_{p}s_{i}\left(N_{D}-N_{D}^{+}\right)I, \end{equation} \begin{equation}\label{e9} \frac{\partial J}{\partial x}=0, \end{equation} where $I_{d} (=\frac{\beta_{T}}{s_{i}})$ is the dark irradiance which is also the homogeneous intensity that controls the conductivity of the crystal. Usually $I$ is such that $N\ll N_{D}$, $N\ll N_{A}$ and $N_{A}\ll N_{D}^{+}$. The space charge field $E_{sc}$ under these approximations turns out to be \begin{equation}\label{e10} E_{sc}=E_{0}\frac{I_{\infty} +I_{D}}{I+I_{D}}+E_{p}\frac{I_{\infty} -I}{I+I_{D}}-\frac{k_{B}T}{e}\frac{1}{I+I_{D}}\frac{\partial I}{\partial x}, \end{equation} where $E_{o}$ and $E_{p}=\frac{k_{p} \gamma_{R} N_{A}}{e\mu}$ are the external bias field to the crystal and photovoltaic field, respectively. In the above derivation we have assumed that the power density of the optical field and space charge field attain asymptotically constant values i.e., $I(x \rightarrow \pm\infty,z)=I_{\infty}$ and $E_{sc}(x \rightarrow \pm\infty,z)=E_{o}$. The refractive index change $\triangle n$ in a noncentrosymmetric PR crystal~\citep{14,17,18,19,20, 21, 22, 23, 24, 33, 34, 35}, such as $BaTiO_{3}$, $LiNbO_{3}$ etc., is due to the linear electro-optic effect (Pockel's) and is given by \begin{equation}\label{e11} \Delta n =-\frac{1}{2}n^{3}r_{e}E_{sc}, \end{equation} \begin{equation}\label{e12} =-\frac{1}{2}n^{3}r_{e}\left[E_{0}\frac{I_{\infty} +I_{D}}{I+I_{D}}+E_{p}\frac{I_{\infty} -I}{I+I_{D}}-\frac{k_{B}T}{e}\frac{1}{I+I_{D}}\frac{\partial I}{\partial x}\right], \end{equation} where $r_{e}$ and $n$ are effective linear electro-optic coefficient and average refractive index, respectively. In centrosymmetric PR materials the index change is due to the quadratic electro-optic response to a photoinduced internal field and can be expressed ~\citep{35,41} as \begin{equation}\label{e13} \Delta n =-\frac{1}{2}n_{b}^{3}g_{e}\epsilon_{0}^{2}(\epsilon_{r}-1)^{2}E_{sc}^{2} , \end{equation} \begin{equation}\label{e14} \triangle n =-\frac{1}{2}n_{b}^{3}g_{e}\epsilon_{0}^{2}(\epsilon_{r}-1)^{2}\left[E_{0}\frac{I_{\infty} +I_{D}}{I+I_{D}}+E_{p}\frac{I_{\infty} -I}{I+I_{D}}-\frac{k_{B}T}{e(I+I_{D})}\frac{\partial I}{\partial x}\right]^{2}, \end{equation} where $g_{e}$, $\epsilon_{r}$, $\epsilon _{0}$ and $n_{b}$ are the effective quadratic electro-optic coefficient, relative dielectric constant, free space permittivity and refractive index of the crystal, respectively; it is assumed that the (dc) polarization is in the linear regime. Depending on the sign of $g_{e}$ the nonlinearity is either self-focusing or defocusing. The nonlinear property of the PR media is evident from equations(12) and (14) since the refractive index is intensity dependent. Different terms in the above expressions are responsible for the existence of different types of solitons. Screening and photovoltaic solitons respectively owe their existence to the first and second term in these expressions. When both first and second terms are dominant, one would expect screening photovoltaic solitons. The third term in each of these equations arises due to the diffusion process and is not in general responsible for the formation of solitons, however, it is responsible for self deflection of solitons ~\citep{14,21,41}. \section{ Spatial Optical Solitons Owing to Single Photon PR Phenomenon} \subsection{Historical Development } Way back in 1992, Segev et al~\citep{17,18} first detected PR solitons in quasi steady-state regime. Soon PR screening solitons were predicted and identified~\citep{19,20,21}. Two different varieties of steady state screening solitons were subsequently investigated and identified in different biased noncentrosymmetric media under varieties of experimental configurations~\citep{19,20,37,38,39,40,42,43,44,45}. Subsequently, screening solitons were also predicted and observed in centrosymmetric materials, particularly in KLTN crystals~\citep{37,41}. Another family of optical solitons were also observed in unbiased PR crystals which exhibits photovoltaic effect. These solitons, popularly known as photovoltaic solitons, have been observed experimentally in 1D as well as 2D configurations~\citep{46,47,48}. Besides these, PR solitons which are combination of the screening and photovoltaic solitons were also predicted and successfully observed~\citep{42,49}. They owe their existence to both photovoltaic effect and spatially nonuniform screening of the applied field, and, are also known as screening photovoltaic (SP) solitons. By controlling the magnitude of bias field they can be converted to screening solitons or photovoltaic solitons.\\\\ PR crystals like $LiNbO_{3}$ possess high second-order susceptibility, hence, can be used for parametric processes such as second-harmonic generation. Recently, it has become very popular, primarily due to its ultra slow PR relaxation time owing to which any written solitonic channels can act as optical waveguides for long time even after turning off the soliton beam. These soliton induced waveguides have been used both for switching devices ~\citep{50} and enhancing the second harmonic conversion efficiency ~\citep{51}. Though non-centrosymmetric PR crystals possess large nonlinearity, their typical response time is in the range of milliseconds, hence, these materials are not suitable for very fast reconfigurable waveguide channels and switches. On the otherhand using paraelectric centrosymmetric PR crystals, particularly KLTN, DelRe et al.~\citep{52} demonstrated dynamic switching using electroholography~\citep{52,53,54}. Spatial solitons are created in paraelectrics that in turn is employed to guide the signal beam of non-photorefractive wavelength. The soliton induced waveguide remains stable even at high power levels, and can be modified via the electro-optic effect by varying external biasing field, ensuring fast change of propagation properties of the signal beam. Fast electro-optic response of $ KLTN$ makes them useful in important applications like optical communications and signal processing and this has paved the way to fast nanosecond applications of PR spatial solitons ~\citep{52,53,54}. \subsection{Modified Nonlinear Schr$\ddot{o}$dinger Equation($\textbf{mNLSE}$)} Most of the solitary wave experiments in PR media employ optical beams with one transverse dimension, therefore, we assume the optical beam is such that no dynamics is involved in y-direction and it is permitted to diffract only along $x$ direction. The external bias field $E_{0}$ and optic axis of the crystal are directed along the $x$ axis.The extraordinary refractive index $\hat{n}_{e}$ in such cases~\citep{14,20,42,55} is given by $({\hat{n}_{e})^{2}}=n_{e}^{2}-n_{e}^{4}r_{e}E_{sc},$ where $n_{e} $ is the unperturbed extraordinary index of refraction. The electric field of the soliton forming beam is assumed to be $\overrightarrow{E}= \overrightarrow{x}\Phi(x,z)exp[i(kz-\omega t)],$ where $k=k_{0}n_{e}$, $k_{0}=\frac{2\pi}{\lambda _{0}}$, $\lambda_{0}$ is the free space wavelength of the optical field. Employing slowly varying envelope approximation for $\Phi $ in the Maxwell's equations we obtain \begin{equation}\label{e15} i\frac{\partial \Phi}{\partial z}+\frac{1}{2k}\frac{\partial^{2}\Phi}{\partial x^{2}}-\frac{1}{2}k_{0}n_{e}^{3}r_{e}E_{sc}\Phi=0. \end{equation} By virtue of the space charge field evaluated in section (3), we obtain following~\citep{14,21,69,70,71,72,73} modified nonlinear Schr\"{o}dinger equation (mNLSE): \begin{equation}\label{e16} i\frac{\partial A}{\partial\xi }+\frac{1}{2}\frac{\partial^{2} A}{\partial s^{2}}-\frac{\beta (1+\rho)A}{1+|A|^{2}}- \frac{\alpha (\rho-|A|^{2})A}{1+|A|^{2}}+\frac{\delta A}{1+|A|^{2}}\frac{\partial |A|^{2}}{\partial s}=0, \end{equation} where $\xi=\frac{z}{k_{0}n_{e}x_{0}^{2}}$, $s=\frac{x}{x_{0}}$, $\alpha=(k_{0}x_{0})^{2}(n_{e}^{4}r_{e}/2)E_{p}$, $\beta=(k_{0}x_{0})^{2}(n_{e}^{4}r_{e}/2)E_{0}$, $\rho=\frac{I_{\infty}}{I_{d}}$, $\delta=(k_{0}^{2}x_{0}r_{e}n_{e}^{4}k_{B}T)/(2e)$, and $A=\sqrt{\frac{n_{e}}{2\eta_{0}I_{d}}}\Phi$. Above mNLSE is the governing equation for varieties of bright, dark and gray solitons. Two parameters $\alpha $ and $\beta$ paly very important role in the formation of these solitons, while $\delta$, which is associated with the diffusion term, is not directly responsible for soliton formation. The diffusion processes is primarily responsible for bending of trajectories of propagating solitons, hence, large value of $\delta$ influences the trajectory of bending. \section{Screening Optical Spatial Solitons} Screening solitons could be created in biased nonphotovoltaic PR crystals (i.e., $\alpha$=0 ), hence, if we neglect diffusion then the equation for screening solitons reduces to \begin{equation}\label{e17} i\frac{\partial A}{\partial\xi }+\frac{1}{2}\frac{\partial^{2} A}{\partial s^{2}}-\beta( 1+\rho)\frac{A}{1+|A|^{2}}=0. \end{equation} \textbf{Bright Screening Solitons:} The soliton forming beam for bright solitons vanishes at infinity, therefore $I_{\infty}=\rho=0$. We express the stationary bright soliton as $A(s,\xi)=p^{\frac{1}{2}}y(s)\exp(i\nu \xi) $, where $\nu$ is the nonlinear shift in propagation constant and y(s) is a normalized real function. In addition, $0\leq y(s)\leq1$, $y(0)=1$, $\dot{y}(0)=0$ and $y(s\rightarrow{\pm \infty})=0$. The parameter \emph{p} represents the ratio of the peak intensity $(I_{max})$ to the dark irradiance $I_{d}$, where $I_{max}=I(s=0)$. Substitution for $A(s,\xi)$ into equation (17) yields \begin{equation}\label{e18} \frac{d^{2}y}{ds^{2}}-2\nu y-2\beta\frac{y}{1+py^{2}}=0. \end{equation} Use of boundary condition and integration of above equation leads to \begin{equation}\label{e19} s=\pm\frac{1}{(2\beta)^{1/2}}\int_{y}^{1}\frac{p^{1/2}}{[\ln(1+p\widehat{y}^{2})-\widehat{y}^{2}\ln(1+p)]^{1/2}} d\widehat{y}. \end{equation} The bright profile $y(s)$ can be obtained using numerical procedure and can be easily shown that these solitons exist when $\beta>0 $ i.e., $E_{0}$ is positive.\\\\ \textbf{Dark Screening Solitons:} In appropriate media, equation (17) also admits dark solitary wave solutions~\citep{14,21} which are embedded in a constant intensity background, therefore, $I_{\infty}$ is finite, hence, $\rho$ is also finite. In addition, they exhibit anisotropic field profiles with respect to $s$. We take following ansatz for stationary solutions: $A(s,\xi)=\rho^{1/2} y(s)\ \exp(i\nu\xi)$, where $\nu$ is the nonlinear shift in propagation constant and $y(s)$ is a normalized real odd function of $s$ satisfying $y(0)=0, y(s\rightarrow\pm \infty)=\pm1, \frac{dy}{ds}=\frac{d^{2}y}{ds^{2}}=0$ as $s\rightarrow \pm\infty$.\ Substituting $A$ in equation (17) we obtain, \begin{equation}\label{e20} \frac{d^{2}y}{ds^{2}}-2\nu y-2\beta ( 1+\rho)\frac{y}{1+\rho y^{2}}=0. \end{equation} By virtue of integration and use of boundary condition, we immediately obtain \begin{equation}\label{e21} s=\pm\frac{1}{(-2\beta)^{1/2}}\int_{y}^{0}\frac{d\widehat{y}}{[(\widehat{y}^{2}-1)-\frac{(1+\rho)}{\rho}\ln\frac{1+\rho \widehat{y}^{2}}{1+\rho}]^{1/2}}. \end{equation} Obviously, these solitons exist only when $\beta<0 $ i.e., $E_{0}$ is negative. Unlike their bright counterpart, these dark screening photovoltaic solitons do not possess bistable property. \\\\ \textbf{Gray Screening Solitons:} Besides bright and dark solitons, equation (17) also admits another interesting class of solitary waves, which are known as gray solitons~\citep{14,21}. In this case too, wave power density attains a constant value at infinity i.e., $I_{\infty}$ is finite, and hence, $\rho$ is finite. To obtain stationary solutions, we assume \begin{equation}\label{e22} A(s,\xi)=\rho^{1/2}y(s)\exp[i(\nu\xi+\int^{s}\frac{Jd\widehat{s}}{y^{2}(\widehat{s})})], \end{equation} where $\nu$ and $J$ are nonlinear shift in propagation constant and a real constant, respectively; $y(s)$ is a normalized real even function of $s$ with properties $y^{2}(0)=m\ (0<m<1)$, $\dot{y}(0)=0$, $y(s\rightarrow \pm\infty)=1$ and all derivatives of $y(s)$ are zero at infinity. The parameter \emph{m} describes grayness, i.e., the intensity $I(0)$ at the beam center is $I(0)=mI_{\infty}$. Substitution of the above ansatz for $A$ in equation (17) yields \begin{equation}\label{e23} \frac{d^{2}y}{ds^{2}}-2\nu y-2\beta ( 1+\rho)\frac{y}{1+\rho y^{2}}-\frac{J^{2}}{y^{3}}=0. \end{equation} The values of $J$ and $\nu$ can be obtained using boundary conditions mentioned earlier. Inserting these values and after integrating once, we get \begin{equation}\label{e24} \left(\frac{dy}{ds}\right)^{2}=2\nu(y^{2}-1)+\frac{2\beta}{\rho}(1+\rho)\ln \left(\frac{1+\rho y^{2}}{1 + \rho }\right)+2(\nu+\beta)\left(\frac{1-y^{2}}{y^{2}}\right). \end{equation} Profiles of dark solitons can be obtained easily by numerical integration of above equation. Unlike bright or dark solitons, the phase of gray solitons is not constant across $s$, instead varies across $s$ and they can exist only when $\beta<1$ and $ m<1$. \section{ Optical Spatial Vector Solitons } In previous sections, our discussions were confined to optical spatial solitons which are solutions of a single dynamical equation. These solutions arise due a single optical beam with specific polarization. However, there are instances where two or more optical beams may mutually get trapped and propagate without distortion. These beams could be of same or different frequencies or polarization. they are mutually trapped and depend on each other in such a way that undistorted propagation of one is sustained by other and vice versa. In order to describe self and mutually trapped propagation of more than one soliton forming optical beams, we need to solve a set of more than one coupled solitary waves. Solutions of this set of coupled NLS equations are called vector solitons if they preserve their shape. \subsection{Incoherently Coupled Spatial Vector Solitons} Steady state incoherently coupled solitons are most extensively studied vector solitons in PR media ~\citep{14,58,59,60,61,62,63,64,65,66,67,68,69,70}. These solitons exist only when the two soliton forming beams possess same polarization and frequency and are mutually incoherent. Four different varieties of solitons i.e., bright-bright, dark-dark, bright-dark and gray-gray~\citep{63,69,70} have been observed. Since two beams are mutually incoherent, no phase matching is required and they experience equal effective electro-optic coefficients. The idea of two incoherently coupled solitons has been generalized and extended to soliton families where number of constituent solitons are more than two. \subsection{Coupled mNLSE Owing to Single Photon Phenomenon} We consider a pair of mutually incoherent optical beams of same frequency and polarization which are propagating in a lossless PR crystal along \emph{z}-direction. The optical c-axis of the crystal and polarization of both beams are oriented along the \emph{x}-direction. These beams are allowed to diffract only along the \emph{x}-direction and \emph{y}-dynamics has been implicitly omitted in the analysis. These beams are expressed as $\overrightarrow{E_{j}} =\overrightarrow{x}\Phi_{j}(x,z)\exp(ikz)$, $j=1, 2 $ and $\Phi_{j}$ is the slowly varying envelope of optical field which satisfies following equation: \begin{equation}\label{e25} i\frac{\partial \Phi_{j}}{\partial z}+\frac{1}{2k}\frac{\partial^{2}\Phi_{j}}{\partial x^{2}}-\frac{k_{0}n_{e}^{3}r_{33}E_{sc}}{2}\Phi_{j}=0, \end{equation} Neglecting diffusion effect, the space charge field can be obtained from equation (10) as \begin{equation}\label{e26} E_{sc}=E_{0}\frac{I_{\infty}+I_{d}}{I+I_{d}}+E_{P}\frac{I_{\infty}-I}{I+I_{d}}, \end{equation} where $I(x,z)=n_{e}/(2\eta_{0} )(|\Phi_{1}|^{2}+|\Phi_{2}|^{2})$ is total power density of two beams, whose value at a distance far away from the center of the crystal is $I_{\infty}=I(x\rightarrow \pm\infty)$. Substituting the expression of $E_{sc}$ in equation (25), we derive following equation: \begin{eqnarray}\label{e27} i\frac{\partial A_{j}}{\partial\xi }+\frac{1}{2}\frac{\partial^{2} A_{j}}{\partial s^{2}}-\beta(1+\rho) \frac{A_{j}}{(1+|A_{j}|^{2}+|A_{3-j}|^{2})}\nonumber \\ -\alpha\frac{(\rho-|A_{j}|^{2}-|A_{3-j}|^{2})A_{j}}{(1+|A_{j}|^{2}+|A_{3-j}|^{2})} =0, \end{eqnarray} where $A_{j}=\sqrt{\frac{n_{e}}{2\eta_{0}I_{d}}}\Phi_{j}$;\ $\alpha$, $\beta$, $\xi$, $s$ and $\rho$ are defined earlier. Above set of two coupled Schr\"{o}dinger equations can be examined for bright-bright, bright-dark, dark-dark, gray-gray screening, photovoltaic as well as screening photovoltaic solitons~\citep{59,60,61,62,63,64,65,66,67,68,69}. In the theoretical front, numerical method to solve above set of coupled equations was developed by Christodoulides et. al.~\citep{58}. Though this method has been used extensively, it fails to identify the existance of large family of solitons. Konar et al~\citep{70} has developmed a method which captures those solitons missed out by Christodoulides et. al.~\citep{58}. In next few lines we describe the method to identify bright-dark solitons only. For elaboration readers are referred to ~\citep{14,59,60,61,62,63,64,65,66,67,68,69,70}.\\\\ \textbf{Bright-dark soliton:} We express $A_{1}=p^{1/2} f(s)\exp(i\mu\xi)$ and $A_{2}=\rho^{1/2} g(s)\exp(i\nu\xi)$, where $f(s)$ and $g(s)$ respectively represents envelope of bright and dark beams. Two positive quantities $p$ and $\rho$ represent the ratios of their maximum power density with respect to the dark irradiance $I_{d}$. Therefore, bright-dark soliton pair obeys following coupled ordinary differential equations: \begin{eqnarray}\label{e28} \frac{d^{2}f}{ds^{2}}-2\left[\mu + \frac{\beta(1+\rho)}{1+pf^{2}+\rho g^{2}}\right]f=0, \end{eqnarray} and \begin{eqnarray}\label{e29} \frac{d^{2}g}{ds^{2}}-2\left[\nu + \frac{\beta(1+\rho)}{1+pf^{2}+\rho g^{2}}\right]g=0. \end{eqnarray} A particular solution of above equations is obtained assuming $f^{2}+g^{2}=1$. Use of boundary conditions gives $\mu=-\frac{\beta}{\Lambda} \ln(1+\Lambda)$ and $\nu=-\beta$, where $ \Lambda= (p-\rho)/(1+\rho)$. When peak intensities of two solitons are approximately equal ($\Lambda<<1$), the soliton solution~\citep{58,71} leads to $A_{1}=p^{1/2}$\ sech$[(\beta\Lambda)^{1/2}s]\exp[-i\beta(1-\Lambda/2)\xi],$ and $A_{2}=\rho^{1/2}\tanh[(\beta\Lambda)^{1/2}s]\exp[-i\beta\xi],$ which exist only when $(\beta\Lambda >0)$. \section{Two-Photon Photorefractive Phenomenon} In previous sections, we have discussed properties of optical spatial solitons which owe their existence due to single photon PR phenomenon. Recently, Ramadan et al.~\citep{72} created bright spatial solitons in a biased $BSO$ crystal using two-step excitation process. Electrons were first excited to the conduction band by a background beam, and then they were excited towards higher levels in the conduction band by a second optical beam of larger wavelength. Using this two-step process in a biased $BSO$ crystal, Ramadan et al.~\citep{72} demonstrated the self-confinement of a red beam at 633 nm supported by another optical beam at 514.5 nm. Recently, Castro-Camus and Magana~\citep{73} also presented an identical model of two-photon PR phenomenon which includes a valance band (VB), a conduction band (CB) and an intermediate allowed level (IL). A gating beam of photon energy $\hbar\omega_{1}$ is used to maintain a quantity of excited electrons from the valance band (VB) to an intermediate allowed level (IL) which are subsequently excited to the conduction band by another signal beam with photon energy $\hbar\omega_{2}$. The signal beam can induce a spatial dependent charge distribution leading to a nonlinear change of refractive index in the medium. Based on Castro-Camus and Magana's model, several authors have investigated two-photon screening and photovoltaic solitons ~\citep{10,14,68,74,75} which owe their existence due to two-photon PR phenomenon. \subsection{Optical Nonlinearity and Evolution Equation of Solitons } In order to estimate optical nonlinearity owing to two-photon PR phenomenon, we need to evaluate the space charge field which can be obtained from the set of rate, current and Poisson's equations proposed by Castro-Camus et al.~\citep{73}. Instead of doing that here we refer interested readers to references~\citep{14,75} and straightaway use the expression of space charge field $E_{sc}$ due to two-photon PR phenomenon~\citep{75}. We assume the soliton forming optical beam of intensity $I_{2}$ is propagating along the $z$ direction of the crystal which is permitted to diffract only along the $x$ direction. The optical beam is polarized along the $x$ axis which is also the direction crystal c-axis and the external bias field. The soliton forming beam is taken as $\overrightarrow{E}=\overrightarrow{x}\Phi(x,z)\exp[i(kz-\omega t)]$, where the symbols have been defined earlier. The crystal is biased with external voltage $V$ and connected with external resistance $R$ and it is illuminated with a gating beam of constant intensity $I_{1}$. We assume that both the power density and space charge field are uniform at large distance from the center of the soliton forming beam, thus, $ I_{2} (x\rightarrow \pm\infty, z)=I_{2\infty}$ =constant and $E_{sc} (x\rightarrow \pm\infty, z)=E_{0}$. Neglecting the effect of diffusion, the space charge field turns out to be \begin{eqnarray}\label{e30} E_{sc}&=&gE_{a}\frac{(I_{2\infty }+I_{2d})(I_{2}+I_{2d}+\frac{\gamma_{1}N_{A}}{s_{2}})}{(I_{2 }+I_{2d})(I_{2\infty}+I_{2d}+\frac{\gamma_{1}N_{A}}{s_{2}})} \nonumber \end{eqnarray} \begin{eqnarray}\label{e30} +E_{p}\frac{s_{2}(gI_{2\infty }-I_{2})(I_{2}+I_{2d}+\frac{\gamma_{1}N_{A}}{s_{2}})}{(I_{2 }+I_{2d})(s_{1}I_{1}+\beta_{1})}, \end{eqnarray} where $N_{A}$ and $N$ are acceptor density and electron density in conduction band, respectively; $\gamma_{R}$, $\gamma_{1}$ and $\gamma_{2}$ are the recombination factor of the conduction to valence band transition, intermediate allowed level to valence band transition and conduction band to intermediate level transition, respectively; $\beta_{1}$ and $\beta_{2}$ are respectively the thermo-ionization probability constant for transitions from valence band to intermediate level and intermediate level to conduction band; $s_{1}$ and $s_{2}$ are photo-excitation crosses. $E_{p}=\kappa_{p} N_{A}\gamma_{R}/e\mu $ is the photovoltaic field, $I_{2d}=\beta_{2}/s_{2}$ is the dark irradiance, $g= 1/(1+q)$, $q= \frac{e\mu N_{\infty}SR}{d}$, $N_{\infty}=N( x\rightarrow \pm\infty )$. In general, $g$ is bounded between $0\leq g\leq 1$. Under short circuit condition $R=0$ and $g=1$, implying the external electric field is totally applied to the crystal. For open circuit condition $R\rightarrow \infty,$ thus, $g=0$ i.e., no bias field is applied to the crystal. Employing equation (30) and following the procedure employed earlier, the nonlinear Schr\"{o}dinger equation for the normalized envelope can be obtained as~\citep{75,77} \begin{eqnarray}\label{e31} i\frac{\partial A}{\partial\xi }+\frac{1}{2}\frac{\partial^{2} A}{\partial s^{2}}-\beta g\frac{(1+\rho)(1+\sigma+|A|^{2})A}{(1+|A|^{2})(1+\sigma+\rho)}\nonumber \\ -\alpha\eta\frac{(g\rho-|A|^{2})(1+\sigma+|A|^{2})A}{1+|A|^{2}}=0, \end{eqnarray} where $\rho= I_{2\infty}/I_{2d}$, $A= \sqrt{\frac{n_{e}}{2\eta_{0}I_{2d}}}\Phi$, $\beta=(k_{0} x_{0} )^{2} (n_{e}^{4} r_{33}/2) E_{a}$, $\eta=\beta_{2}/(s_{1} I_{1}+\beta_{1}), \sigma=\frac{\gamma_{1}N_{A}}{s_{2}I_{2d}}=\frac{\gamma_{1}N_{A}}{\beta_{2}}$. Equation (31) can be employed to investigate screening, photovoltaic and screening photovoltaic solitons under appropriate experimental configuration. \\\\ \textbf{Bright Screening Solitons:} These solitons, for which $\alpha =0$ and $\rho =0 $, have been studied by several authors~\citep{14,73,74,75,76,77}. In the low amplitude limit i.e., when $|A|^{2}<<1$, equation (31) reduces to \begin{equation}\label{e32} i\frac{\partial A}{\partial\xi }+\frac{1}{2}\frac{\partial^{2} A}{\partial s^{2}}-\frac{\beta}{1+\sigma}\left(1+\sigma-\sigma|A|^{2}\right)A =0. \end{equation} The one-soliton solution of above equation is given by \begin{equation}\label{e36} A(s,\xi)=p^{1/2}sech\left[\left(\frac{\beta p \sigma}{1+\sigma}\right)^{1/2}s\right]\\exp\left[i\frac{\beta(p\sigma-2\sigma-2)}{2(1+\sigma)}\xi\right]. \end{equation} \\\ \textbf{Dark screening solitons:} For this case $\rho \neq0 $, thus when $|A|^{2}<<1$, \ equation (31) reduces to \begin{equation}\label{e37} i\frac{\partial A}{\partial \xi}+\frac{1}{2}\frac{\partial^2 A}{\partial s^2}-\frac{\beta(1+\rho)}{(1+\sigma+\rho)}(1+\sigma-\sigma |A|^2)A =0. \end{equation} The dark soliton solution of above equation turns out to be \begin{equation}\label{e38} A(s,\xi)=\rho^{1/2}\tanh\left[-\left(\frac{\beta\rho\sigma}{1+\sigma+\rho}\right)^{1/2}s\right]exp\left[\frac{i\beta(1+\rho)(\rho\sigma-\sigma-1)\xi}{1+\sigma+\rho}\right]. \end{equation} These solitons could be observed in $SBN$ since, they have an intermediate level required for two step excitation. In addition to bright and dark solitons, equation (31) also predicts steady state gray solitons which were investigated by Zhang et. al.~\citep{76}. Characteristics of these solitons are similar to one-photon PR gray spatial solitons. For example, they require bias field in opposite to the optical c-axis and their FWHM is inversely proportional to the square root of the absolute value of the bias field. Proceeding in a similar way we can study bright and dark photovoltaic solitons. \section{ Modulation Instability(MI)} Modulation instability (MI) is an inherent characteristic of wave propagation in nonlinear media. It refers to unstable propagation of a continuous or quasi-continuous wave(CW) in such a way that the wave disintegrates into large number of localized coherent structures after propagating some distance through nonlinear physical media. MI occurs as a result of interplay between nonlinearity and dispersion in temporal domain, and between nonlinearity and diffraction in the spatial domain~\citep{2,5,81,82,83,84,85,86,87,88,89,90}. A continuous wave (CW) or quasi-continuous wave radiation propagating in a nonlinear medium may suffer instability with respect to weak periodic modulation of the steady state and results in the breakup of CW into a train of ultra short pulses. In spatial domain, for a narrow beam, self phase modulation exactly balances the diffraction and a robust spatial soliton is obtained, while a broad optical beam disintegrates into many filaments during propagation in the same self-focusing nonlinear medium.\\\\\ MI has been made extensively studied in a wide range of physical systems like fluids~\citep{79}, plasmas ~\citep{81}, Bose Einstein condensates~\citep{82}, discrete nonlinear systems~\citep{83}, negative index materials~\citep{84}, soft matter~\citep{85}, PR media~\citep{86,87}, optical fibers ~\citep{88} etc. MI typically occurs in the same parameter region where solitons are observed. In fact, the filaments that emerge from the MI process are actually trains of almost ideal solitons. Therefore, the phenomenon of MI can be considered as a precursor of soliton formation and has been found in both coherent beams and incoherent beams. MI has been extensively investigated in PR media ~\citep{80,90,91}. Single as well as two-photon PR media have been considered to analyse instability characteristics. Moreover, not only noncentrosymmetric but centrosymmetric PR media have been examined as well ~\citep{90}. Unlike noncentrosymmetric media, in a centrosymmetric media the characteristics of this instability is independent of the external applied field. In the next section, we examine the MI of a broad optical beam in a biased two-photon non-centrosymmetric photovoltaic PR medium. \subsection{ MI Gain under Linear Stability Framework} In order to find out MI gain, we consider an optical beam with large transverse spatial dimension. Since we are confining our present interest on the stability of a broad bright beam of finite transverse extension, therefore $I_{2\infty}=0$ and hence, $\rho=0$. Therefore, the evolution equation of the broad optical field reduces to \begin{eqnarray}\label{e36} i\frac{\partial A}{\partial\xi }+\frac{1}{2}\frac{\partial^{2} A}{\partial s^{2}}-\beta g\frac{(1+\sigma+|A|^{2})}{(1+|A|^{2})(1+\sigma)}A\nonumber \\ +\alpha\eta\frac{|A|^{2}(1+\sigma+|A|^{2})}{(1+|A|^{2})}A=0. \end{eqnarray} Equation (36) admits a steady state CW solution $A(\xi,s)=\sqrt{P}\exp[i\Psi(\xi)],$ where $P$ is the initial input power at $\xi=0$ and $\Psi(\xi)$ is the nonlinear phase shift which increases with propagation distance $\xi$ according to \begin{eqnarray}\label{e37} \Psi(\xi)=-g\beta\frac{(1+\sigma+P)}{(1+P)(1+\sigma)}\xi+\alpha\eta\frac{P(1+\sigma+P)}{1+P}. \end{eqnarray} The initial stage of MI can be investigated in the linear stability framework, under which the stability of the steady state solution is examined by introducing a perturbation in the amplitude of the beam envelope so that the perturbed field now becomes: \begin{eqnarray}\label{e38} A(\xi,s)=\left[\sqrt{P}+a(\xi,s)\right]\exp[i\Psi(\xi)], \end{eqnarray} where $a(\xi,s)$ is an arbitrarily small complex perturbation field such that $a(\xi,s)<<\sqrt{P}$. Substituting the perturbed field in equation (36) and retaining terms linear only in the perturbed quantity, the evolution equation for the perturbation field is obtained as \begin{equation}\label{e39} i\frac{\partial a}{\partial \xi}+\frac{1}{2}\frac{\partial ^{2}a}{\partial s^{2}}+\alpha\eta P(a+a^{*})+\frac{\alpha\eta\sigma P(a+a^{*})}{(P+1)^{2}}+\frac{g\beta\sigma P(a+a^{*})}{(1+\sigma)(1+P)^{2}}=0 \end{equation} where $*$ denotes complex conjugate. The spatial perturbation $a(\xi,s)$ is assumed to be composed of two sideband plane waves \begin{equation}\label{e40} a(\xi,s)=u\cos(K\xi-\Omega s)+iv \sin (K\xi-\Omega s), \end{equation} where $u $ and $v$ are the real amplitudes of the perturbing field, $K$ and $\Omega$ being the wave number and spatial frequency of the perturbations, respectively. Substitution for the perturbation field into its evolution equation yields: \begin{equation}\label{e41} \begin{pmatrix} \Pi^{-} & -K\\ K & \Pi^{+}\\ \end{pmatrix} \begin{pmatrix} u\\ v\\ \end{pmatrix} = 0\\, \end{equation} where $\Pi^{-}=-\Pi^{+}+\frac{2g\beta\sigma P}{(1+\sigma)(1+P)^{2}}+\frac{2\alpha\eta\sigma P}{(1+P)^{2}}+2\alpha\eta P$ and $\Pi^{+}=\Omega^{2}/2$. Equation(40) possesses a nontrivial solution only when the following dispersion relation holds good: \begin{equation}\label{e42} K^{2}=-\frac{\Omega^{2}}{2}\left(\frac{2g\beta\sigma P}{(1+\sigma)(1+P)^{2}}+\frac{2\alpha\eta\sigma P}{(1+P)^{2}}+2\alpha\eta P-\Omega^{2}/2\right). \end{equation} If the wave number of the perturbation becomes complex, then the instability will set in with exponential growth of the perturbation field $a(\xi,s)$ resulting in the filamentation of the broad beam into a number of filaments. Thus, the propagating broad optical beam will be unstable. The growth rate $g(\Omega) (=2Im(K))$ of the MI is obtained as \begin{equation}\label{e43} g(\Omega)=\sqrt{2}\Omega\left(\frac{2g\beta\sigma P}{(1+\sigma)(1+P)^{2}}+\frac{2\alpha\eta\sigma P}{(1+P)^{2}}+2\alpha\eta P-\Omega^{2}/2\right)^{1/2}. \end{equation} \subsection{Gain Spectrum of Instability } MI gain is achievable in a photovoltaic PR (PVPR) crystal for a given range of frequency and specified range of values of $g,\sigma,\eta,\alpha,\beta$ and $P$. Parameters $\eta$ and $\sigma$ are always positive, while $\alpha$ and $\beta$ can be both positive or negative depending on the media and the polarity of the external bias field. In an unbiased PVPR media $( E_{a}=0\ i.e.,\beta=0)$ , thus K is always positive if $\alpha< 0$. Hence, usually in such media, modulation instability cannot set in. However, in such media with the application of external field of appropriate magnitude and polarity MI can set in and grow as long as \begin{equation}\label{e44} \frac{2g\beta\sigma P}{(1+\sigma)(1+P)^{2}}> 2|\alpha|\eta P\left(1+\frac{\sigma}{(1+P)^{2}}\right)+\frac{\Omega^{2}}{2}. \end{equation} Therefore, with the application of external electric field, it is possible to initiate modulation instability and control the growth of the instability in those media where MI was hitherto prohibited. On the other hand, if for a given value of beam power MI growth rate is finite in an unbiased PVPR medium with positive value of $\alpha$ , then the instability growth rate can be enhanced or decreased with the application of external electric field of appropriate magnitude and polarity. Or in other words, growth rate of the instability can be controlled with the application of external field of chosen polarity and magnitude. MI can takes place only below the critical frequency $\Omega_{c}$ which is given by \begin{equation}\label{e45} \Omega_{c}=\pm\left[\frac{g\beta\sigma P}{(1+\sigma)(1+P)^{2}}+\frac{\alpha\eta\sigma P}{(1+P)^{2}}+\alpha\eta P\right]^{1/2}. \end{equation} The instability is most efficient and reaches its maximum at $g=g_{m}$ when $\Omega=\Omega_{m}$, where $g_{m}= \frac{\Omega_{c}^{2}}{2}$ and $\Omega_{m}=\frac{\Omega_{c}^{2}}{\sqrt{2}}$. To examine the MI growth, we take a typical Cu:KNSBN crystal. Intermediate energy level is included in Cu:KNSBN crystal and photovoltaic field is in the direction of optic axis. At $\lambda_{0}=0.5\mu m$, crystal parameters are $n_{e}=2.27,\ E_{p}=2.8\times10^{6} Vm^{-1}$, $\eta=1.5\times10^{-4},\sigma=10^{4}$, $r_{33}=200\times10^{-12}m/V$. The scaling parameter $x_{0}=10\mu m$, $\alpha=117.3$ and $g=1$. We take three different values of $E_{a}$, in particular, $E_{a}=-2\times 10^{6} Vm^{-1}$, $0$ and $2\times 10^{6} Vm^{-1}$ which corresponds to $\beta=-83.79,0$ and $ 83.79$ respectively. Figure (1) depicts the MI gain spectrum $g(\Omega)$ as a function of perturbation frequency $\Omega$. The variation of $\Omega_{m}$ with $P$ for three values of $\beta$ has been depicted in figure (2). Initially $\Omega_{m}$ increases with the increase in the value of $P$ then decreases with the increase in $P$. In order to examine the role played by $\alpha$ on the growth of the instability, we have demonstrated in figure (3) the variation of maximum growth $g_{m}$ with the normalized beam power for three different value of $\alpha$. As expected, a higher value of $\alpha$ enhances the growth of the instability. In conclusion, with the application of external electric field, it is possible to initiate modulation instability and control the growth of the instability in those media where MI was hitherto prohibited. \section{Conclusion } A brief review of some selected developments in the field of optical spatial solitons in PR media has been presented. Underlying mechanism responsible for the formation of solitons have been discussed for both single and two-photon PR media. Vector solitons, particularly, incoherently coupled solitons due to single photon and two-photon PR phenomena have been highlighted. Existence of some missing solitons pointed out. Modulation instability which is a precursor to soliton formation has been also considered. Important applications of PR solitons have been highlighted. \section*{Acknowledgment} This work is supported by SAP programme of University Grants Commsion (UGC), Government of India. One of the authors SK would like to thank UGC for this.\\\ \newpage
\section{Introduction} \nin The $\{\beta\}$-expansion approach, discussed here, was originally proposed in \cite{SM}. The aim was to construct generalizations of the BLM approach \cite{BLM} at the levels higher than the NNLO one while the first method to fix the BLM-type scale for the RG-invariant quantities was developed in \cite{GK}. This $\{\beta\}$-expansion was used to explore multiple power $\beta$-function generalization of the Crewther relation in the $\MSbar$-scheme for the nonsinglet (NS) corrections to the Adler $D$-function and to the Bjorken sum rule of the polarized lepton-nucleon scattering \cite{KM1}. Expanding this form of the generalized Crewther relation in powers of $\alpha_s$ and keeping the single power of the QCD $\beta$-function only, one can recover the generalized Crewther relation with the single $\beta$-function factor. The existence of this $\MSbar$-scheme relation was discovered at the $\alpha_s^3$-level \cite{Broadhurst:1993ru} and confirmed later on in \cite{Baikov:2010je} at the $\alpha_s^4$ order. This relation follows from the consideration of the AVV quark current triangle diagram not only in the massless quark-parton model \cite{Crewther:1972kn}, which respects conformal symmetry, but in the case, when the insertion of higher-order QCD corrections to this triangle diagram are also taken into account \cite{Gabadadze:1995ei}. Theoretical validity of the generalized $\MSbar$-scheme Crewther relation, presented as the additional term with factored out single power of the $\beta$-function was studied in \cite{Crewther:1997ux,Braun:2003rp}, where its validity in all orders of perturbation theory was investigated. More recently the $\{\beta\}$-expansion approach was explored in \cite{KM2} in relation to its analog, used in \cite{PMC1,PMC2,PMC3,PMC4,PMC5} for various applications of the Principle of Maximal Conformality (PMC) proposed in \cite{PMC}. Note that the main aim of PMC, which is similar to the seBLM method in \cite{SM}, is to construct a new high-order representation of the BLM approach by absorbing all terms proportional to the $\beta$-function coefficients into the scales of each integer power of the coupling $\alpha_s$ in perturbative series for the RG-invariant quantities. For the NS Adler function and the Bjorken polarized sum rule the coefficients of these modified series should respect the relations, which follow from the conformal symmetry and the original Crewther relation of \cite{Crewther:1972kn} (for the recent theoretical studies of the consequences of the conformal symmetry in QED and QCD see \cite{Kataev:2013vua}). \section{Comparison of the complete and incomplete $\{\beta\}$- expansions for the $D^{NS}$-function} \nin Following the work \cite{KM2}, let us clarify first the differences between the complete and unique $\{\beta\}$-expansion \cite{SM} and the incomplete one used in the studies of \cite{PMC1,PMC2,PMC3,PMC4,PMC5}. Within the complete $\{\beta\}$-expansion the expression for the perturbative coefficients of the N$^3$LO approximation of the $D^{NS}$-function \beq \label{Dns} D^{NS}(a_s)=1+\sum_{n=1}^{n=4} d_{n} a_s^{n} \eeq is expressed through the coefficients of the $\beta$-function of the colour $SU(N_c)$ gauge group model \beq \label{bf} \beta(a_s)=\mu^2\frac{\partial a_s}{\partial \mu^2}= -\sum_{i\geq 0} \beta_i(N_F)a_s^{i+2}~~. \eeq in the following form: \bea \label{d1b} d_1&=&d_1[0], \\ \label{d2b} d_2&=&\beta_0(N_F)d_2[1]+d_2[0], \\ \nonumber d_3&=&\beta_0^2(N_F)d_3[2]+\beta_1(N_F)d_3[0,1] \\ \label{d3b} &&+ \underline{\beta_0(N_F)d_3[1]}+d_3[0], \\ \nonumber d_4&=&\beta_0^3(N_F)d_4[3]+\beta_1(N_F)\beta_0(N_F)d_4[1,1 ] \\ \nonumber &&+\beta_2(N_F)d_4[0,0,1]+ \beta_0^2d_4[2] \\ \label{d4b} &&+ \underline{\beta_1(N_F)d_4[0,1]}+\underline{\beta_0(N_F)d_4[1]}+d_4[0], \eea where $N_F$ is the number of fermion flavours and the {\it underlined} terms were neglected in similar expansions used in \cite{PMC1,PMC2,PMC3,PMC4,PMC5}. The reason of neglecting them is related to the fact that the authors of these works define their $\{\beta\}$-expansions from the traditional expressions f0r $d_i$ coefficients expanded in powers of $N_F$, namely \bea \label{1} d_1&=&N_F^0d_1 \\ \label{2} d_2&=&N_Fd_{21}+d_{20} \\ \label{3} d_3&=&N_F^2d_{32}+N_Fd_{31}+d_{30} \\ \label{4} d_4&=&N_F^3d_{43}+N_F^2d_{42}+N_Fd_{41}+d_{40}~~~. \eea However, it is already known that to formulate the generalized BLM approach at the NNLO using Eqs.(7-9), it is necessary to take into account some extra information \cite{GK,SM}. Within the approach of \cite{SM} extra terms, which allow one to obtain the complete $\{\beta\}$ expansion of $d_3$ in Eq.(5), are the analytical contributions of the multiplet of light gluinos $n_{\tilde{g}}$ to the $O(\alpha_s^3)$ approximations of the $D^{NS}$-function \cite{Chetyrkin:1996ez} and to the $\beta$-function of the $SU(N_c)$ group, which was evaluated in the $\MSbar$-scheme at the three-loop level in \cite{Clavelli:1996pz}. The application of new degrees of freedom $n_{\tilde{g}}$ in both $D^{NS}(a_s)$ and $\beta(a_s)$-functions at the $O(\alpha_s^3)$ level allowed splitting in the $\{\beta\}$-expansion of the $\beta_1(N_F)$ and $\beta_0(N_F)$-dependent terms, both contributing to the $N_F$-term of $d_3$ in Eq.(\ref{3}). We have thus obtained \cite{SM} the elements in the RHS of Eqs.(\ref{d1b}-\ref{d3b}), which define the following new matrix representation for $D^{NS}$: \beq \label{matrix} D^{NS}(Q^2)=1+\sum_{n\geq 1}\sum_{l}a_s^{n}(Q^2)d_{n}[l]B_l(N_F) \eeq In Eq.(\ref{matrix}), the $B_l(N_F)$ factors are the products of the $\beta$-function coefficients of Eqs.(\ref{d1b}-\ref{d4b}), $d_n(N_F)=d_{n}[l]B_l(N_F)$ are the $N_F$-dependent coefficients in Eqs.(\ref{1}-\ref{4}) while the elements $d_{n}[l]$ do not depend on the numbers of flavours $N_F$. Note that in view of the absence of an analytical result for the gluino contributions to $D^{NS}$ at the $O(\alpha_s^4)$ level, we are unable to get most part of the terms in the $\{\beta\}$-expansion of $d_4$ in Eq.(\ref{d4b}). Indeed, only the leading $\beta_0^{3}(N_F)d_4[3]$-contribution is known from analytical calculations of \cite{Broadhurst:1993ru}. This information was already used in the all-order generalization of the BLM approach of \cite{Beneke:1994qe}, based on absorbing into the BLM scale these renormalon-type terms only. Since we are interested in the resummation of all $\{\beta\}$-dependent terms, we will consider only certain expressions at the $O(\alpha_s^3)$-level. In the $\MSbar$-scheme, at this order of perturbation theory the elements of the $\{\beta\}$-expansion for $D^{NS}$ have the following analytic form \cite{SM}: \bea \label{d10a} d_1[0]&=&\frac{3}{4}~C_F \\ \label{d21a} d_2[1]&=&\bigg(\frac{33}{8}-3\zeta_3\bigg)~C_F \\ \label{d20a} d_2[0]&=&-\frac{3}{32}~C_F^2+\frac{1}{16}~C_FC_A \\ \label{d32a} d_3[2]&=&\bigg(\frac{151}{6}-19\zeta_3\bigg)~C_F \\ \label{d301a} d_3[0,1]&=&\bigg(\frac{101}{16}-6\zeta_3\bigg)~C_F \\ \label{d31a} d_3[1]&=&\bigg(-\frac{27}{8}-\frac{39}{4}\zeta_3+15\zeta_5\bigg)~ C_F^2 \\ \nonumber &&-\bigg(\frac{9}{64}-5\zeta_3+\frac{5}{2}\zeta_5\bigg)~C_FC_A \\ \label{d30a} d_3[0]&=&-\frac{69}{128}~C_F^3+\frac{71}{64}~C_F^2C_A \\ \nonumber &&+ \bigg(\frac{523}{768}-\frac{27}{8}\zeta_3\bigg)~C_FC_A^2 \eea Note once more that in the PMC studies of \cite{PMC1,PMC2,PMC3,PMC4,PMC5} an analog of the $d_3[1]$-term was absent ( or nullified). Therefore, the remaining $\MSbar$-scheme contributions to Eq.(\ref{d3b}) will differ from the ones presented in Eqs.(\ref{d301a}) and (\ref{d30a}). \section{ The $\bar{MS}$ -scheme generalized Crewther relation and the $\{ \beta \}$-expansion for the Bjorken polarized sum rule } \nin The Bjorken polarized sum rule, which is still interesting for phenomenological studies \cite{Khandramai:2011zd,Deur:2014vea}, is defined as \beq S_{Bjp}=\int_0^1 g_1^{lp-ln}(x,Q^2)dx= \frac{g_A}{6}C^{Bjp}(a_s) . \eeq Its coefficient function $C^{Bjp}$ contains the NS and singlet (SI) contributions \beq C^{Bjp}(a_s)=C^{Bjp}_{NS}(a_s)+C^{Bjp}_{SI}(a_s) . \eeq The existence of the SI term at the $O(\alpha_s^4)$ level was demonstrated in \cite{Larin:2013yba}, though its analytical expression is not yet fixed by direct diagram-by-diagram calculations. The $\{\beta\}$-expansion pattern is now applied to the coefficients $c_{n}$ of the perturbative approximation for $C^{Bjp}_{NS}(a_s)$ \cite{Baikov:2010je}: \beq C^{Bjp}_{NS}(a_s)=1+\sum_{n=1}^{n=4} c_n a_s^{n}\,, \eeq \bea \label{c1b} c_1&=&c_1[0] \\ \label{c2b} c_2&=&\beta_0(N_F)c_2[1]+c_2[0] \\ \nonumber c_3&=&\beta_0^2(N_F)c_3[2]+\beta_1(N_F)c_3[0,1] \\ \label{c3b} &&+ \underline{\beta_0(N_F)c_3[1]}+c_3[0] \\ \nonumber c_4&=&\beta_0^3(N_F)c_4[3]+\beta_1(N_F)\beta_0(N_F)c_4[1,1 ] \\ \nonumber &&\!\!\!\!+\beta_2(N_F)c_4[0,0,1]+ \beta_0^2c_4[2] \\ \label{c4b} &&\!\!\!\!+ \underline{\beta_1(N_F)c_4[0,1]}+\underline{\beta_0(N_F)c_4[1]}+c_4[0]. \eea These coefficients are related to similar ones, which enter into the $\{\beta\}$-expansion of the perturbative series for $D^{NS}$ through the multiple power $\beta$-function form of the generalization of the Crewther relation \cite{KM1}. Note that the application of the $\MSbar$-scheme generalization of the Crewther relation, considered in \cite{PMC4}, gives the $\{\beta\}$- expanded expressions for the coefficients of the $C^{Bjp}_{NS}(a_s)$-series without the terms underlined in Eqs.(\ref{c3b},\ref{c4b}). We will show that the absence of these terms in the studies of \cite{PMC1,PMC2,PMC3,PMC4,PMC5} {\it contradicts} the existing analytical $\MSbar$-scheme $O(\alpha_s^3)$ results for the $D^{NS}(Q^2)$ function \cite{Gorishnii:1990vf,Surguladze:1990tg,Chetyrkin:1996ez} and the $\MSbar$-scheme generalization of the Crewther relation \cite{Broadhurst:1993ru,Baikov:2010je} written down in the multiple power $\beta$-function representation of \cite{KM1}. This part of the talk follows from the studies of \cite{KM2}. The $\MSbar$-scheme single $\beta$-function expression for the generalized Crewther relation has the following form: \beq \label{Crew1} D^{NS}(a_s)C^{Bjp}_{NS}(a_s)=1+\frac{\beta(a_s)}{a_s}K(a_s)~~. \eeq Here $K(a_s)=K_1a_s+K_2a_s^2+K_3a_s^3 +O(a_s^4)$ is the polynomial, where the known coefficient $K_1$ depends on the $SU(N_c)$ Casimir operator $C_F$ while the coefficients $K_2$ and $K_3$ also known analytically depend on $C_F$, $C_A$, $T_F$ and $N_F$. This form, originally discovered in \cite{Broadhurst:1993ru} at the $O(a_s^3)$ level, was recently confirmed by direct $O(a_s^4)$ calculations of $D^{NS}$ and $C^{Bjp}_{NS}$ performed in the colour $SU(N_c)$ gauge group theory \cite{Baikov:2010je}. In \cite{KM1}, it was demonstrated that Eq.(\ref{Crew1}) can be rewritten as \bea \label{CrwPol} \!\!\!\!D^{NS}(a_s)C^{Bjp}_{NS}(a_s)\!\!=\!\! 1+\frac{\beta(a_s)}{a_s}\sum_{n=1}^{3} \bigg(\frac{\beta(a_s)}{a_s}\bigg)^{n-1}\!\!\!\!\!\! P_n(a_s)&& \\ =1+\sum_{n\geq 1}\sum_{r\geq 1}P_n^{r}[k,m]C_F^{k}C_A^{m}a_s^{r}~~~&&\nonumber \eea where $k+m=r$ and the coefficients $P_n^{r}[r,m]$ contain rational fractions and Riemann $\zeta$-functions of odd arguments. In Eq.(\ref{CrwPol}), the known coefficients of the polynomials $P_n(a_s)$ \textit{ do not depend} on $T_F N_F$ (for a more obvious clarification of this property see the second expression in Eq.(\ref{CrwPol})) and are expressed by means of the coefficients of the $\{\beta\}$-expansion as \bea P_1(a_s)&=&-a_s\bigg[c_2[1]+d_2[1]\nonumber \\ &&+ a_s\bigg( c_3[1]+d_3[1]+d_1(c_2[1]-d_2[1])\bigg) \nonumber \\ &&+a_s^2\bigg(c_4[1]+d_4[1]+d_1(c_3[1]+d_3[1])\bigg)\nonumber \\ &&+d_2[0]c_2[1]+d_2[1]c_2[0]\bigg] \label{P1}\\ \nonumber P_2(a_s)&=&a_s\bigg[c_3[2]+d_3[2]+a_s\Big( c_4[2]+d_4[2] \\ &&-d_1(c_3[2]-d_3[2])\Big)\bigg] \label{P2} \\ P_3(a_s)&=&-a_s\bigg[c_4[3]+d_4[3]\bigg]\label{P3} \eea Using Eq.(\ref{CrwPol}) the following relations between the elements of the $\{\beta\}$-expansions of Eqs.(\ref{d1b}-\ref{d4b}) and Eqs.(\ref{c1b}-\ref{c4b}) were obtained \cite{KM1}: \bea &&0=c_n[0]+d_n[0]+\sum_{l=1}^{n-1}d_l[0]c_{n-l}[0] \label{Crew} \\ &&C_F\bigg(-\frac{21}{8}+3\zeta_3\bigg) = -c_2[1]-d_2[1]= \nonumber \\ &&-c_3[0,1]-d_3[0,1]=\ldots\label{P11} \\ \label{P21} &&\!\!\!\!\!\!-c_3[1]-d_3[1]-d_1(c_2[1]-d_2[1])= \\ \nonumber &&\!\!\!\!\!\! -c_4[0,1]-d_4[0,1]-d_1(c_3[0,1]-d_3[0,1])=\ldots \eea Using Eqs.(\ref{d10a}-\ref{d30a}) for $d_n[l]$ and solving then either Eq.(\ref{Crew}) (initial Crewther relation \cite{Crewther:1972kn}) or Eq.(\ref{P11}) (from \cite{KM1}) we got the elements of the $\{\beta\}$-expansion for $C^{Bjp}_{NS}$ at the $O(a_s^3)$ level: \bea \label{c10a} c_1[0]&=&-\frac{3}{4}~C_F \\ \label{c21a} c_2[1]&=&-\frac{3}{2}~C_F \\ \label{c20a} c_2[0]&=&\frac{21}{32}~C_F^2-\frac{1}{16}~C_FC_A \\ \label{c32a} c_3[2]&=&-\frac{151}{24}~C_F \\ \label{c301a} c_3[0,1]&=&\bigg(-\frac{59}{16}+3\zeta_3\bigg)~C_F \\ \label{c31a} c_3[1]&=&\bigg(\frac{83}{24}-\zeta_3\bigg)~ C_F^2 \\ \nonumber &&+\bigg(\frac{215}{64}-6\zeta_3+\frac{5}{2}\zeta_5\bigg)~C_FC_A \\ \label{c30a} c_3[0]&=&-\frac{3}{128}~C_F^3-\frac{65}{64}~C_F^2C_A \\ \nonumber &&- \bigg(\frac{523}{768}-\frac{27}{8}\zeta_3\bigg)~C_FC_A^2~~. \eea Apart from the presented above analytical expressions, we come to the definite theoretical conclusions. First, we note that Eq.(\ref{Crew}) is the consequence of the Crewther relation \cite{Crewther:1972kn} and of the conformal symmetry. However, it does not give us a possibility to say anything about the {\it scheme-independence} of the coefficients $d_i[0]$ and $c_j[0]$ even within the MS-like schemes. We can only conclude that the coefficients of the PMC series are {\it scheme-dependent} but obey {\it the scheme-independent} relation, which {\it follows from the conformal symmetry}. Second, the chain of Eqs.(\ref{P21}) clearly demonstrates, that the $\MSbar$ analytical calculations of $D^{NS}$, $C_{Bjp}$ and the $\MSbar$-scheme generalizations of the Crewther relations of Eq.(\ref{Crew1}), (\ref{CrwPol}) do not allow one to {\it neglect (or nullify)} the terms $d_3[1]$, $d_4[0,1]$ in the $\{\beta\}$-expansion of the coefficients Eq.(\ref{d3b}) and Eq.(\ref{d4b}) of the $D^{NS}$ RG-invariant function and of their analogs $c_3[1]$, $c_4[0,1]$ in the perturbative expansion of the Bjorken polarized sum rule. Indeed, their absence contradicts the analytical result on the LHS of Eq.(\ref{P21}), obtained in \cite{KM1} from the $\MSbar$-scheme generalization of the Crewther relation. In view of this, the theoretical and phenomenological studies of the works \cite{PMC1,PMC2,PMC3,PMC4,PMC5}, where the discussed above nonzero terms were neglected, should be reconsidered. This was done in part in \cite{KM2} and we will summarize below the concrete foundations of this work. \section{The definition of the scale-fixing prescription } \nin To define the generalized BLM approach within the {\it complete} and {\it unique} $\{\beta\}$-expansion approach of \cite{SM}, one should absorb all $\beta$-dependent terms of the $\{\beta\}$-expanded coefficients into the scales of the coupling constants. Following the study of \cite{KM2}, let us absorb all $\beta$-dependent terms of the coefficients in Eqs.(\ref{d2b},\ref{d3b}) into the new scales of the related perturbative expansions of the $D^{NS}$-function and $C^{Bjp}_{NS}$ RG-invariant quantities. Using the solution of the RG-equation in Eq.(\ref{bf}) we reexpress the QCD running coupling constant $a_s(\mu^2)$ in terms of the new one $a_s^{'}=a_s(\mu^{'2})$ in the following form considered in \cite{KM2}, namely: \bea \label{redefin} a_s(t)&=&a^{'}-\beta(a_s^{'})\frac{\Delta}{1!} +\beta(a_s^{'})\partial_{a_s^{'}}\beta(a_s^{'})\frac{\Delta^2}{2!} \\ \nonumber &+&\beta(a_s^{'})\partial_{a_s^{'}}(\beta(a_s^{'})\partial_{a_s^{'}}\beta(a_s^{'})) \frac{\Delta^3}{3!}+\dots \eea The term $\Delta$ defines the shift of the scales as \beq \Delta=t-t^{'}=ln(\mu^2/\mu^{'2}) \eeq where $t=ln(Q^2/\mu^2)$ and $t^{'}=ln(Q^2/\mu^{'2})$. To define all-order generalization of the BLM approach proposed in \cite{SM}, it is necessary to introduce the coupling constant dependent shift \beq \label{shift} \Delta=\Delta(a_s^{'})=\Delta_0+a_s^{'}\beta_0\Delta_1+(a_s^{'}\beta_0)^2\Delta_2+ \dots, \eeq where the coupling constant dependence of this shift was first introduced in \cite{GK} in the process of the first formulation of the NNLO generalization of the BLM approach. Fixing now $Q^2=\mu^{'2}$ we obtain the $\{\beta\}$-expansions of the transformed to the new scale coefficients $d_n^{'}$ of the perturbative expressions for the $D^{NS}$-function. They have the following form \cite{KM2}: \bea \label{d1bp} d_1^{'}&=&d_1[0] \\ d_2^{'}&=&\beta_0\,d_2[1]+ d_2[0] -\beta_0\Delta_{0} \label{d2bp}\\ d_3^{'}&=&\beta_0^2 (d_3[2]- 2d_2[1]\Delta_{0} + \Delta_{0}^2) \nonumber\\ &&\!\!\!\!+ \beta_1(d_3[0,1]-\Delta_{0}) +\beta_0(d_3[1] -2d_2[0]\Delta_{0}) \nonumber \\ &&\!\!\!\!+d_3[0]+\beta_0^2\Delta_{1} \label{d3bp} \eea For the sake of generality, we also present the expression for the fourth term, which due to still incomplete analytical information on its $\beta$-expansion can not be involved in the concrete numerical studies \bea d_4^{'}&=&\beta_0^3(d_4[3] -3d_3[2]\Delta_{0}+ 3d_2[1]\Delta_{0}^2 -\Delta_{0}^3 \nonumber\\ \nonumber &&- 2(\Delta_{0}-d_2[1])\Delta_{1}) \\ \nonumber &&+\beta_1\beta_0(d_4[1,1]-(3d_3[0,1]+2d_2[1])\Delta_{0} \\ \nonumber &&+\frac{5}2\Delta_{0}^2-\Delta_{1}) +\beta_2(d_4[0,0,1]- \Delta_{0}) \\ \nonumber &&+\beta_0^2(d_4[2] -3d_3[1]\Delta_{0} \\ \nonumber &&+3d_2[0]\Delta_{0}^2-2d_2[0]\Delta_{1}) \\ \nonumber &&+\beta_1(d_4[0,1]-2d_2[0]\Delta_{0}) \\ \nonumber &&+\beta_0(d_4[1]-3d_3[0]\Delta_{0}) \\ &&+ d_4[0]-\beta_0^3\Delta_{2}\label{d4bp} \eea The general idea of \cite{KM2} is to absorb all $\{\beta\}$-dependent terms in Eqs.(\ref{d2bp}-\ref{d4bp}), including the ones omitted in \cite{PMC1,PMC2,PMC3,PMC4,PMC5}, namely the terms proportional to $d_3[1]$, $d_4[0,1]$ and $d_4[1]$. Then, we accumulate these terms in ``shift'' coefficients $\Delta_0$, $\Delta_1$ and $\Delta_2$, which defines the new BLM (PMC-type) scales of $D^{NS}(a'_s)$. We will present here the results of application of this procedure at the $O(a_s^3)$ level only, where all coefficients of the $\beta$-dependent terms are already determined, see Eqs.(\ref{d21a}-\ref{d31a}) and Eqs.(\ref{c21a}-\ref{c31a}). \section{The concrete analytical and numerical $O(\alpha_s^3)$ studies} \nin Following the studies \cite{KM2}, and solving Eqs.(\ref{d2bp},\ref{d3bp}) with respect to $\Delta_{0},\Delta_{1}$ and similar expressions for $C^{Bjp}_{NS}$ in the case of ordinary QCD, we arrive at the concrete analytical and numerical results for the parameters in the defined in Eq.(\ref{shift}) scale $\Delta$ of the PMC-type BLM generalization of $O(a_s^3)$ approximations for $D^{NS}$ and $C^{Bjp}_{NS}$: \bea \label{DeltaD0} \Delta_0&=&d_2[1]= \frac{11}{2}-4\zeta_3=0.69177\\ \label{DeltaC0} \bar{\Delta}_0&=&c_2[1]=-2 \\ \label{DD1n} \Delta_1&=&\frac{1}{\beta_0^2}\big[\beta_0^2(d_3[2]-d_2[1]^2) \nonumber \\ &&+\beta_1(d_3[0,1]-d_2[1]) \nonumber\\ &&+\beta_0(d_3[1]-2d_2[0]d_2[1])\big] \label{CD1n}\\ \bar{\Delta}_1&=&\frac{1}{\beta_0^2}\big[\beta_0^2(c_3[2]-c_2[1]^2) \nonumber \\ &&+\beta_1(c_3[0,1]-c_2[1]) \nonumber\\ &&+\beta_0(c_3[1]-2c_2[0]c_2[1])\big] \,. \eea Eqs.(\ref{DD1n},\ref{CD1n}) contain noticeable contributions of the terms omitted in \cite{PMC1,PMC2,PMC3,PMC4,PMC5}, that are proportional to $d_3[1]$ and $c_3[1]$. Note that for normalization, used here, we have $a_s=\alpha/\pi$, $\beta_0=11/4-N_F/6$ and $\beta_1= 51/8-19N_F/24$. The approximate numerical expressions for the coefficients of the $\{\beta\}$-expansion for the $D^{NS}$ and $C^{Bjp}_{NS}$ RG-invariant functions read \bea \label{d10n} d_1[0]&=&1~~~~~~~~~~~~~~~~~~~c_1[0]=-1 \\ d_2[1]& \approx &0.69~~~~~~~~~~c_2[1]=-2 \\ d_2[0]&\approx &0.083~~~~~~~~~~c_2[0]\approx0.917 \\ d_3[2]&\approx &3.105~~~~~~~~~~c_3[2] \approx -6.39\\ d_3[0,1]&\approx&-1.2~~~~~~~c_3[0,1]\approx -0.108 \\ d_3[1]&\approx&13.926~~~~~~~~~c_3[1]\approx -10 \\ d_3[0]&\approx&-35.87~~~~~~~c_3[0] \approx 35.03 \eea We note poor convergence of the $O(\alpha_s^3)$ approximations of the perturbative series constructed from the respective conformal symmetry coefficients $d_n[0]$ and $c_n[0]$. Indeed, the concrete result for the normalized $NS$ contribution to the $e^+e^-$ R-ratio, which is related to the $D^{NS}(Q^2)$-function, has the following form \cite{KM2}: \bea \label{DNSI} R^{NS}(s)&=&1+a_s(s_{\rm PMC})+0.0833~a_s^2(s_{\rm PMC}) \\ \nonumber &-& 35.872~a_s^3(s_{\rm PMC})+O(a_s^4) \eea where the scale is defined through the solution of Eq.(\ref{DeltaD0}) and Eq.(\ref{DD1n}) . We will present it for $N_F=3$ numbers of active flavours with $\beta_0=2.25$ and $\beta_1=4$. It has the following expression: \beq \label{QD} s_{\rm PMC}=s \cdot \exp[-0.69-3.98\beta_0a_s^{'}(s)] \eeq Note that at the NLO we reproduce the standard BLM coefficient, which is rather small. However, the value of the NNLO coefficient is negative and huge. A similar feature was already observed in the case of applications of the first generalization of the BLM approach based on resummation of the $N_F$-dependent corrections \cite{GK}. This result of \cite{GK} was confirmed in \cite{PMC1}. Applying the same procedure to $C^{Bjp}_{NS}$ in \cite{KM2} we got \bea C^{Bjp}_{NS}(Q^2)\!\!\!\!\!&=&\!\!\!\!\!1-a_s(Q^2_{\rm PMC})+0.917 a_s(Q^2_{\rm PMC}) \\ \nonumber &&+ 35.03 a_s^2(Q^2_{\rm PMC})+O(a_s^3) \eea where \beq Q^2_{\rm PMC}=Q^2\cdot \exp[-2-7.32\beta_0a_s^{'}(Q^2)] \eeq Similar results were previously obtained at the NNLO for the Bjorken polarized sum rule within the procedure of \cite{GK} in \cite{Kataev:1992jm}. \section{Conclusion} \nin We would like to emphasize that the proposed in \cite{SM} and used later on in \cite{KM1,KM2} $\{\beta\}$-expansion approach allows one to fix the special terms $d_3[0]$ and $c_3[0]$ of the $\MSbar$-scheme series for the $e^+e^{-}$ characteristic $R^{NS}(s)$ and for the Bjorken polarized sum rule. They satisfy the relations, which follow from the {\it conformal symmetry}. However, leaving only these terms in the $O(a_s^3)$ approximations for the special generalizations of the BLM procedure one gets huge coefficients related with $O(a_s^3)$ level. In view of this, the direct applications of theoretically interesting PMC-type (or seBLM-type) approximations in the phenomenological studies should be treated with care. \section*{Acknowledgements} \nin The work was supported in part by the Russian Foundation for Basic Research, Grant No. 14-01-00647. The work of MS was also supported by the BelRFFI--JINR, grant F14D-007. One of us (ALK) wishes to thank S.Narison for invitation to the QCD-2014 and hospitality in Montpelier. At the final stage of this work, the studies of the consequences of the multiple $\beta$-function representation of the generalized Crewther relation by ALK were supoported in part by RSCF, Grant N 14-22-00161. \input{bib_sample} \end{document}
\section{Introduction} The concept of an efficient dominating set in a graph was introduced by Biggs~\cite{Biggs} as a generalization of the notion of a perfect error-correcting code in coding theory. Given a (simple, finite, undirected) graph $G=(V,E)$, we say that a vertex \emph{dominates} itself and each of its neighbors. An \emph{efficient dominating set} in $G$ is a subset of vertices $D\subseteq V$ such that every vertex $v \in V$ is dominated by precisely one vertex from $D$. Efficient domination has several interesting applications in coding theory and resource allocation of parallel processing systems~\cite{Biggs, LLT97, LS90}. The notion of an efficient dominating set appeared in the literature under various other names such as: \emph{perfect code}, \emph{$1$-perfect code}, \emph{independent perfect dominating set}, and \emph{perfect dominating set}. Note, however, that the name \emph{perfect dominating set} has also been used in the literature to denote a subset of vertices $D\subseteq V$ such that every vertex $v \in V\setminus D$ is dominated by precisely one vertex from $D$. See~\cite{LT02} for a nice historical overview of the notion of efficient dominating set. A graph is \emph{efficiently dominatable} if it contains an efficient dominating set. All paths are efficiently dominatable, and a cycle $C_k$ on $k$ vertices is efficiently dominatable if and only if $k$ is a multiple of $3$. Bange \emph{et al.}~\cite{BBS88} showed that if a graph $G$ has an efficient dominating set, then all efficient dominating sets of $G$ have the same cardinality, which equals the minimum cardinality of a dominating set of $G$. The efficient domination (ED) problem consists in determining whether the input graph is efficiently dominatable. The ED problem is \textsf{NP}-complete even for restricted graph classes such as planar cubic graphs~\cite{Kratochvil91}, bipartite graphs~\cite{YL96}, planar bipartite graphs~\cite{LT02}, chordal bipartite graphs~\cite{LT02}, chordal graphs~\cite{YL96}, and line graphs of planar bipartite graphs of maximum degree three~\cite{BHN}. On the other hand, the ED problem is polynomial for several graph classes, including trees~\cite{BBS88,FH91}, block graphs~\cite{YL96}, interval graphs~\cite{CL94, CRC95, KE00, KMM95}, circular-arc graphs~\cite{CL94,KE00}, cocomparability graphs~\cite{C97,CRC95}, bipartite permutation graphs~\cite{LT02}, permutation graphs~\cite{LLT97}, distance-hereditary graphs~\cite{LT02}, trapezoid graphs~\cite{LLT97,L98}, split graphs~\cite{CL93}, dually chordal graphs~\cite{BLR}, AT-free graphs~\cite{BLR,BKM99}, and hereditary efficiently dominatable graphs~\cite{Milanic}. The efficient domination problem has also been studied from a parameterized point of view, see, e.g.,~\cite{Cesati,Guo-Niedermeier}. In this paper, we consider two weighted versions of the ED problem. In its decision form, the minimization version of problems can be stated as follows: \medskip \begin{center} \fbox{\parbox{0.85\linewidth}{\noindent {\sc Minimum Weight Efficient Dominating Set (Min-WED)}\\[.8ex] \begin{tabular*}{.93\textwidth}{rl} \emph{Input:} & A graph $G$, vertex weights $w:V\to \mathbb{Z}$, an integer $k$.\\ \emph{Question:} & Does $G$ contain an efficient dominating set $D$ \\ & of total weight $w(D) := \sum_{x\in D}w(x)\le k$? \end{tabular*} }} \end{center} \medskip The maximization version of problem ({\sc Maximum Weight Efficient Dominating Set} \hbox{\sc (Max-WED)}), can be defined analogously, replacing the condition $w(D) \le k$ with $w(D) \ge k$. Clearly, a graph $G=(V,E)$ contains an efficient dominating set if and only if $(G,w,|V|)$ is a yes instance to the {\sc{Min-WED }} problem, where $w(x) = 1$ for all $x\in V$. Consequently, the {\sc{Min-WED }} problem is \textsf{NP}-complete in every class of graphs where the ED problem is \textsf{NP}-complete. On the other hand, the {\sc{Min-WED }} problem is solvable in polynomial time for trees~\cite{Y92}, cocomparability graphs~\cite{C97,CRC95}, split graphs~\cite{CL93}, interval graphs~\cite{CL94, CRC95}, circular-arc graphs~\cite{CL94}, permutation graphs~\cite{LLT97}, trapezoid graphs~\cite{LLT97,L98}, bipartite permutation graphs~\cite{LT02}, convex bipartite graphs~\cite{BLR} and their superclass interval bigraphs~\cite{BLR}, distance-hereditary graphs~\cite{LT02}, block graphs~\cite{YL96} and hereditary efficiently dominatable graphs~\cite{CMR00,Milanic}. Since negative weights are allowed, the {\sc{Max-WED }} problem is equivalent to the {\sc{Min-WED }} problem. We develop a framework for solving the {\sc{Min-WED }} and {\sc{Max-WED }} problems based on a reduction to the {\sc Maximum Weight Independent Set} problem in the square of the input graph (this is done in Section~\ref{sec:wed-mwis}). We then apply this framework, together with some existing results from the literature, to derive new polynomial cases of the {\sc{Min-WED }} problems, namely the classes of dually chordal graphs and AT-free graphs (in Sections~\ref{sec:dually-chordal} and~\ref{sec:AT-free}, respectively). The class of dually chordal graphs contains the class of strongly chordal graphs, for which the existence of a polynomial-time algorithm for the {\sc{Min-WED }} problem was posed as an open problem by Lu and Tang in~\cite{LT02}. We give a linear-time algorithm for the {\sc{Min-WED }} and {\sc{Max-WED }} problems in the class of dually chordal graphs. Our algorithm for the {\sc{Min-WED }} problem in the class of AT-free graphs is of complexity ${\mathcal O}\big(\min\{nm+n^2,n^\omega\}\big)$, where $\omega < 2.3727$ is the matrix multiplication exponent~\cite{Williams}, and $n$ and $m$ denote the number of vertices and edges of the input graph, respectively. This improves on the existing polynomial-time algorithms for the ED problem in AT-free graphs~\hbox{\cite{BLR,BKM99}}, both of which run in time ${\mathcal O}(n^4)$. \bigskip In Fig.~\ref{fig:classes} below, we show the Hasse diagram of the poset of most of the graph classes mentioned above, ordered with respect to inclusion. For each class, we state the complexity of the {\sc{Min-WED }} problem, denoting by \textsf{NP}-c the fact that the problem is \textsf{NP}-complete in the corresponding class, while in the case of polynomial-time solvability, we state the running times of the fastest known algorithms. The inclusion relations in the figure were verified with help of the Information System on Graph Classes and their Inclusions~\cite{graphclasses}. \begin{figure}[ht] \begin{center} \includegraphics[width=\linewidth]{classes} \caption{ Algorithmic complexity of the weighted efficient domination problem in various graph classes. In the time complexities marked by $^*$, it is assumed that a graph is given by a trapezoid diagram (in the case of trapezoid graphs), by a permutation (in the case of permutation and bipartite permutation graphs), or by a one-vertex-extension ordering (in the case of distance-hereditary graphs). }\label{fig:classes} \end{center} \end{figure} \section{Preliminaries} We only consider finite, simple and undirected graphs. As usual, the \emph{neighborhood} of a vertex $v$ in a graph $G = (V,E)$ is denoted by $N_G(v) := \{u\in V\mid uv\in E\}$ (or simply $N(v)$, if the graph is clear from the context), and the closed neighborhood of $v$ is $N_G[v]:= N_G(v)\cup \{v\}$ (or simply $N[v]$). The \emph{degree} of a vertex $v$ in a graph $G$ is $\deg_G(v) := |N_G(v)|$. The \emph{square} of a graph $G = (V,E)$ is the graph $G^2 = (V,E^2)$ such that $uv\in E^2$ if and only if either $uv\in E$ or $u$ and $v$ are distinct and have a common neighbor in $G$. The \emph{complement} of a graph $G = (V,E)$ is the graph $\overline{G}= (V,\overline{E})$ such that two distinct vertices $u$ and $v$ of $G$ are adjacent in $\overline{G}$ if and only if they are non-adjacent in $G$. An \emph{independent set} in a graph is a set of pairwise non-adjacent vertices, and a \emph{clique} is a set of pairwise adjacent vertices. Given a graph $G$ and a total ordering $\sigma=(v_1,\ldots, v_n)$ of its vertex set, we will denote by $G_{\sigma,i}$ the subgraph of $G$ induced by $\{v_i,v_{i+1},\ldots, v_n\}$, and by $N_{\sigma,i}(v)$ (resp.~$N_{\sigma,i}[v]$), the neighborhood (resp., the closed neighborhood) of a vertex $v\in V(G_{\sigma,i})$ in $G_{\sigma,i}$. \emph{Chordal graphs} are graphs in which every cycle on at least $4$ vertices has a chord (an edge connecting two non-consecutive vertices on the cycle). It is well known that chordal graphs are precisely the graphs that admit a perfect elimination ordering. A \emph{perfect elimination ordering} of a graph $G$ is a total ordering $\sigma = (v_1,\ldots, v_n)$ of its vertex set such that for every $i\in \{1,\ldots, n\}$, vertex $v_i$ is simplicial in $G_{\sigma,i}$, that is, the set $N_{\sigma,i}(v_i)$ is a clique in $G_{\sigma,i}$ (equivalently: in $G$). \emph{Dually chordal graphs} were introduced in~\cite{BDCV} as graphs that admit a certain vertex ordering called a maximum neighborhood ordering. Given two vertices $u$ and $v$ in a graph $G = (V,E)$, vertex $u\in N[v]$ is a \emph{maximum neighbor} of vertex $v$ if $N[w] \subseteq N[u]$ holds for all $w\in N[v]$. A linear ordering $\sigma=(v_1,\ldots, v_n)$ of $V$ is a \emph{maximum neighborhood ordering} of $G$ if for all $i\in \{1,\ldots, n\}$, vertex $v_i$ has a maximum neighbor in the graph $G_{\sigma,i}$. Dually chordal graphs admit several equivalent characterizations and form a generalization of \emph{strongly chordal graphs}, a well-known subclass of chordal graphs properly containing the classes of trees and interval graphs. In fact, strongly chordal graphs are exactly the hereditary dually chordal graphs, that is, graphs for which each induced subgraph is a dually chordal graph. Algorithmic aspects of dually chordal graphs were treated systematically in~\cite{BCD98}. A partial order may be viewed as a transitive directed acyclic graph. A \emph{comparability graph} is the graph obtained by ignoring the edge directions of a transitive directed acyclic graph. A graph is \emph{cocomparability} if its complement is a comparability graph. An \emph{asteroidal triple} in a graph $G$ is a subset $I = \{u,v,w\}$ of three pairwise non-adjacent vertices such that for every vertex $x\in I$, the two vertices in $I\setminus \{x\}$ are contained in the same connected component of the graph $G-N[x]$. A graph $G$ is said to be \emph{AT-free}~\cite{COS97} if it contains no asteroidal triples. AT-free graphs form a large class of graphs containing cographs, interval graphs, permutation graphs, trapezoid graphs, and cocomparability graphs. For further details on graph classes, see~\cite{BLV99,Gol04}. \section{The Reduction} \label{sec:wed-mwis} The decision version of the {\sc{MWIS }} problem can be stated as follows: \medskip \begin{center} \fbox{\parbox{0.85\linewidth}{\noindent {\sc Maximum Weight Independent Set (MWIS)}\\[.8ex] \begin{tabular*}{.93\textwidth}{rl} \emph{Input:} & A tuple $(G,w,k)$ consisting of a graph $G = (V,E)$, \\ & vertex weights $w:V\to \mathbb{Z}$, and an integer $k\in \mathbb{Z}$.\\ \emph{Question:} & Does $G$ contain an independent set $I\subseteq V$ such that $w(I)\ge k$? \end{tabular*} }} \end{center} \medskip This problem, together with its unweighted version, is a classical \textsf{NP}-complete problem, which remains hard even for several restricted graph classes such as triangle-free graphs~\cite{Pol74}, $K_{1,4}$-free graphs \cite{Min80} and planar graphs of maximum degree at most three~\cite{GJ77}. On the other hand, the MWIS problem has been shown to admit polynomial solutions in several graph classes, such as claw-free graphs~\cite{Min80,NT01,OPS08}, and their generalization fork-free graphs~\cite{LM08}, perfect graphs~\cite{GLS84}, and AT-free graphs~\cite{BKM99}. The following simple but useful observation from~\cite{BLR} establishes a connection between efficient dominating sets in a graph $G$ and independent sets in its square. \begin{observation}[\cite{BLR}]\label{lem:eds-wis} Let $G=(V,E)$ be a graph, and let $w$ be a vertex weight function for $G$ defined by $w(x) = |N_G[x]|$ for all $x\in V$. Then, the following statements are equivalent for every $I\subseteq V$: \begin{enumerate} \item[(i)] $I$ is an efficient dominating set in $G$. \item[(ii)] $I$ is a maximum-weight independent set in $G^2$ such that $w(I) = |V|$. \end{enumerate} \end{observation} The above observation (as well as a slightly more general observation from~\cite{Milanic}) has an immediate algorithmic corollary, reducing the ED problem to the MWIS problem in the square of the input graph. In~\cite{BLR,Milanic}, the exact time complexity of such a reduction was not specified; we do this in the following proposition. Given a graph $G$, we denote by $|G|$ its encoding length. \begin{proposition}[\cite{BLR,Milanic}]\label{prop:1} Let $\cal C$ be a graph class for which the {\sc{MWIS }} problem is solvable in time $T(|G|)$ on squares of graphs from $\cal C$. Then, the ED problem for ${\cal C}$ is solvable in time \hbox{${\mathcal O}\big(\min\{nm+n,n^\omega\}+T(|G^2|))$}. \end{proposition} \begin{proof} It is easy to see that the square $G^2$ can be computed in time ${\mathcal O}(n^{\omega})$ using matrix multiplication. Alternatively, assuming that $G$ is given by adjacency lists, computing $G^2$ can be done in time ${\mathcal O}(nm{+n})$, as follows. First, we fix an ordering $\sigma = (v_1,\ldots, v_n)$ of $V$ and order the adjacency lists with respect to $\sigma$. This can be done in linear time, see, e.g.,~\cite[p.~$36$]{Gol04}. Then, we create a new copy of these lists. For every vertex $v\in V$, we process its neighbors in order and for each of them, say $w$, we merge the current adjacency list of $v$ with the adjacency list of $w$ (removing duplicates). Since the lists are ordered, each such merging step can be done in ${\mathcal O}(n)$ time. Hence, the overall time complexity of computing $G^2$ is \[{\mathcal O}(m+n)+\sum_{v\in V}\sum_{w\in N_G(v)}{\mathcal O}(n) = {\mathcal O}(m+n)+{\mathcal O}(n)\cdot\sum_{v\in V}\deg_G(v) {= {\mathcal O}(nm+n+m)} = {\mathcal O}(nm+n)\,,\] where the last equality holds since $n\ge 1$. This justifies the time complexities given in Proposition~\ref{prop:1}. \end{proof} We extend the approach of Brandst\"adt {\it et al.}~\cite{BLR} and Milani\v c~\cite{Milanic} to the weighted versions of the ED problem based on a suggestion made in~\cite{Leitert}. We have the following \begin{theorem}\label{thm:wed-mwis} Let $\cal C$ be a graph class for which the {\sc{MWIS }} problem is solvable in time $T(|G|)$ on squares of graphs from $\cal C$. Then, the {\sc{Min-WED }} and {\sc{Max-WED }} problems are solvable on graphs in ${\cal C}$ in time ${\mathcal O}\big(\min\{nm+n,n^\omega\}+T(|G^2|))$. \end{theorem} The proof of this theorem will rely on two auxiliary lemmas. The first one shows how the problem can be reduced to the case of non-negative weights. \begin{lemma}\label{lem:wed-mwis} Let $(G = (V,E), w, k)$ be an instance to the {\sc{Min-WED }} problem such that $\mu = \min\left\{\min\{w(x)\mid x\in V\},k\right\}<0$. Suppose that $G$ has an efficient dominating set, and let $\gamma$ be the common cardinality of all efficient dominating sets of $G$. For all $x\in V$, let $\tilde w(x) = w(x)+|\mu|$, and let $\tilde k = k+|\mu|\cdot\gamma$. Then, $(G, w, k)$ is a yes instance to the {\sc{Min-WED }} problem if and only if $(G, \tilde w, \tilde k)$ is a yes instance to the {\sc{Min-WED }} problem. In addition, $\tilde \mu = \min\left\{\min\{\tilde w(x)\mid x\in V\},\tilde k\right\}\ge 0$. \end{lemma} \begin{proof} Recall that if a graph $G$ has an efficient dominating set, then all efficient dominating sets of $G$ have the same cardinality, which equals the minimum cardinality of a dominating set of $G$. This implies that if we replace each $w(x)$ with $\tilde w(x) = w(x)+|\mu|$, then the weight of each efficient dominating set will increase by exactly $|\mu|\gamma$, where $\gamma$ is the common cardinality of all efficient dominating sets of $G$. Consequently, $(G,w,k)$ is a yes instance if and only if $(G,\tilde w,k+|\mu|\gamma)$ is a yes instance. The fact that $\tilde \mu\ge 0$ is clear. \end{proof} The second lemma deals with the case of non-negative weights. \begin{lemma}\label{lem:wed-mwis-2} Let $(G = (V,E), w, k)$ be an instance to the {\sc{Min-WED }} problem such that $0 \leq \min\{ \min\{ w(x) \mid x \in V \}, k \}$. Let $M = \max\{\sum_{x\in V}w(x),k\}+1$, let $w'(x) = M\cdot |N_G[x]| - w(x)$, for all $x\in V$, and $k' = M\cdot |V|-k$. Then, $(G, w, k)$ is a yes instance to the {\sc{Min-WED }} problem if and only if $(G^2, w', k')$ is a yes instance to the {\sc{MWIS }} problem. \end{lemma} \begin{proof} On the one hand, if $D$ is an efficient dominating set in $G$ such that $w(D)\le k$, then, by Observation~\ref{lem:eds-wis}, $D$ is an independent set in $G^2$ such that $\sum_{x\in D}|N_G[x]| = |V|$. Therefore, \[w'(D) = M\cdot \sum_{x\in D}|N_G[x]| - \sum_{x\in D}w(x)\ge M\cdot |V|-k = k'\,,\] and $(G^2, w', k')$ is a yes instance to the {\sc{MWIS }} problem. On the other hand, let $I$ be an independent set in $G^2$ with $k'\le w'(I)$. We claim that $\sum_{x\in I}|N_G[x]| = |V|$ and $w(I)\le k$. Since $I$ is an independent set in $G^2$, it is an independent set in $G$ such that the closed neighborhoods (in $G$) of its elements are pairwise disjoint. Therefore, $\sum_{x\in I}|N_G[x]| \le |V|\,.$ Conversely, $k'\leq w'(I)$ implies $\sum_{x\in I}|N_G[x]|\geq \frac{k'}{M}=|V|-\frac{k}{M}>|V|-1,$ since $k<M$. We thus have $\sum_{x\in I}|N_G[x]| = |V|$. By Observation~\ref{lem:eds-wis}, $I$ is an efficient dominating set in $G$. The inequality $w'(I)\ge k'$ is equivalent to \[ M\cdot \sum_{x\in I}|N_G[x]| - \sum_{x\in I}w(x) \ge M\cdot |V|-k\,,\] which implies $w(I) = \sum_{x\in I}w(x)\le k$. Thus, $(G, w, k)$ is a yes instance to the {\sc{Min-WED }} problem, as claimed. \end{proof} \begin{proof}[Proof of Theorem~\ref{thm:wed-mwis}] Consider an instance $(G = (V,E), w, k)$ with $|V| = n$ and $|E| = m$ to the {\sc{Min-WED }} problem. Let $\mu = \min\left\{\min\{w(x)\mid x\in V\},k\right\}$. If $\mu <0$, we transform the instance, using Lemma~\ref{lem:wed-mwis} to an equivalent instance $(G, \tilde w, \tilde k)$ to the {\sc{Min-WED }} problem with non-negative weights and $\tilde k$. The time complexity of this step is dominated by solving the ED problem in $G$, which, by Proposition~\ref{prop:1}, is of the order \hbox{${\mathcal O}\big(\min\{nm+n,n^\omega\}+T(|G^2|))$}. To solve the {\sc{Min-WED }} problem on $(G, \tilde w, \tilde k)$, we apply Lemma~\ref{lem:wed-mwis-2}. The corresponding instance $(G^2, w', k')$ to the {\sc{MWIS }} problem can be computed in time $O(n+m)$. Summarizing, in order to solve the {\sc{Min-WED }} problem, we only need to compute $G^2$ and solve at most two instances of the {\sc{MWIS }} problem on $G^2$. The {\sc{Max-WED }} problem can be reduced in time ${\mathcal O}(n+m)$ to the {\sc{Min-WED }} problem, since a tuple $(G,w,k)$ is a yes instance to the {\sc{Max-WED }} problem if and only if $(G,-w,-k)$ is a yes instance to the {\sc{Min-WED }} problem. Thus the result follows. \end{proof} \section{New Polynomial Cases of the Weighted Efficient Domination Problems} In this section, we exploit the approach developed in Section~\ref{sec:wed-mwis} and develop new polynomial results for the {\sc{Min-WED }} and {\sc{Max-WED }} problems. \subsection{Dually chordal graphs}\label{sec:dually-chordal} In~\cite{LT02}, Lu and Tang wrote that {\it ``$(\ldots)$ it would be of interest to know whether or not there is a polynomial-time algorithm to solve the weighted efficient domination problem on} ($\ldots$) {\it strongly chordal graphs.''} Recently, Brandst\"adt \emph{et al.}~\cite{BLR} gave a linear-time algorithm for the efficient domination problem in the class of dually chordal graphs. The algorithm is a modification of Frank's algorithm (Algorithm~\ref{algo:mwisChordal} below), which solves the {\sc{MWIS }} problem for chordal graphs in linear time~\cite{Frank}. The modification allows to find a maximum weight independent set of~$G^2$ in linear time if $G$ is dually chordal, without computing its square. \begin{algorithm} \caption{\cite{Frank} Algorithm to find a maximum weight independent set in chordal graphs.}\label{algo:mwisChordal} \KwIn{A chordal graph $G=(V,E)$ with $|V| = n$ and a vertex weight function~$\omega$.} \KwOut{A maximum weight independent set $I$ of $G$.} Find a perfect elimination ordering $\sigma = (v_1,\ldots,v_n)$ of $G$ and set $I \leftarrow \emptyset$. \label{line:mwisChorFindPEO} \For{$i = 1$ \KwTo $n$} { If $\omega(v_i) > 0$, mark $v$ and set $\omega(u) \leftarrow \max(\omega(u) - \omega(v_i), 0)$ for all vertices $u \in N_{\sigma,i}(v_i)$. } \For{$i= n~{\bf downto}~1$} { If $v_i$ is marked, set $I \leftarrow I \cup \{v_i\}$ and unmark all $u \in N_{\{v_1,\ldots, v_{i-1}\}}(v_i)$. } \Return{$I$} \end{algorithm} We will make a similar modification of Frank's algorithm so that the obtained algorithm (Algorithm~\ref{algo:EDdc} below) solves the {\sc{Min-WED }} problem for dually chordal graphs. The modification contains two major changes. First, we add a preprocessing step, transforming the given {\sc{Min-WED }} instance (with possibly negative weights) to an instance of the {\sc{MWIS }} problem. Second, we rewrite Algorithm~\ref{algo:mwisChordal} to find a maximum weight independent set of the square of the given graph. By Lemma~\ref{lem:wed-mwis-2}, this will be a solution to the {\sc{Min-WED }} problem. For the first step, we use Lemma~\ref{lem:wed-mwis}. For the second, we need the two following lemmas. \begin{lemma}[\cite{BCD98}]\label{lem:MNO_linear} A maximum neighborhood ordering of $G$ which simultaneously is a perfect elimination ordering of $G^2$ can be found in linear time. \end{lemma} Additionally to finding a maximum neighborhood ordering $\sigma=(v_1,\ldots,v_n)$, the algorithm in \cite{BCD98} also computes a maximum neighbor $m_i$ for each vertex $v_i$ such that for all $i < n$ no vertex $v_i$ is its own maximum neighbor ($v_i \neq m_i$). This is necessary for the next lemma. \begin{lemma}\label{lem:ijInE_iff_mjInE2} Let $G=(V,E)$ be a graph with $G^2=(V,E^2)$ and a maximum neighborhood ordering $\sigma=(v_1, \ldots, v_n)$ where for all $1\le i< n$, $m_i$ is a maximum neighbor of $v_i$ with $v_i \neq m_i$. If $1 \leq i < j \leq n$ and $m_i \neq v_j$, then $v_iv_j \in E^2 \Leftrightarrow m_iv_j \in E$. \end{lemma} \begin{proof} $(\Leftarrow):$ Vertex $v_j$ is adjacent in $G$ to $m_i$ ($m_iv_j \in E$). Thus, the distance between $v_i$ and $v_j$ is at most $2$. So $v_i$ and $v_j$ are also adjacent in $G^2$ ($v_iv_j \in E^2$). $(\Rightarrow):$ Vertices $v_i$ and $v_j$ are adjacent in $G^2$ ($v_iv_j \in E^2$). If $v_iv_j \in E$, then $m_iv_j\in E$. Now assume that $v_iv_j \notin E$. Then vertices $v_i$ and $v_j$ have a common neighbor $v_k$ in $G$, choose the rightmost such vertex (that is, the one maximizing the value of $k$). We distinguish between two cases: \begin{enumerate}[(i)] \item $i<k$. If $m_i= v_k$, then $m_iv_j \in E$ so we may assume that $m_i\neq v_k$. By definition $m_i$ is adjacent to all neighbors of $v_k$ in $G_{\sigma,i}$. This includes $v_j$. \item $k<i$. In this case any maximum neighbor $m_k$ of $v_k$ satisfies $m_k\neq v_i$ (since $v_kv_j\in E$ but $v_iv_j\not \in E$), $v_im_k \in E$ and $m_kv_j \in E$. In particular, $m_k$ is a common neighbor of $v_i$ and $v_j$. Since $m_k = v_p$ for some $p>k$, this contradicts the choice of $v_k$. \end{enumerate} \end{proof} \begin{algorithm}[ht] \label{algo:EDdc} \caption{A linear time algorithm for the {\sc{Min-WED }} problem in dually chordal graphs} \KwIn{An instance $(G,w,k)$ of the {\sc{Min-WED }} problem where $G=(V,E)$ is dually chordal graph.} \KwOut{An efficient dominating set $D$ in $G$ with $w(D)\le k$, if one exists, {\sc No}, otherwise.} \BlankLine { $\mu \leftarrow \min\left\{\min\{w(v)\mid v\in V\},k\right\}$ \nllabel{line:posWeightsStart} \If{$\mu < 0$} { Solve the efficient domination problem on $G$.\nllabel{line:compgamma} \If{$G$ has no efficient dominating set} { \Return{\sc No}; } \Else { Let $\gamma$ be the common cardinality of all efficient dominating sets of $G$. } For all $v \in V$ set $w(v) \leftarrow w(v) + |\mu|$ $k \leftarrow k + |\mu| \cdot \gamma$ \nllabel{line:posWeightsEnd} } $D \leftarrow \emptyset$, $M \leftarrow \max\{\sum_{x\in V}w(x),k\}+1$, $k'\leftarrow M \cdot |V| - k$. \nllabel{line:compMk} \ForAll{$v \in V$\nllabel{line2} } { Set $\omega(v) \leftarrow M \cdot |N[v]| - w(v)$ and $\omega_p(v) \leftarrow 0$.\nllabel{line:compOmega} Set $v$ unmarked and not blocked.\nllabel{line4} } Find a maximum neighborhood ordering $\sigma=(v_1,\ldots,v_n)$ of $G$ which simultaneously is a perfect elimination ordering of $G^2$, with the corresponding maximum neighbors $(m_1, \ldots, m_n)$ where $v_i \neq m_i$ for $1 \leq i < n$. \nllabel{line:compMNO} \For{$i = 1$ \KwTo $n$\nllabel{line6}} { For all $u \in N_{\sigma,i}[v_i]$ set $\omega(v_i) \leftarrow \omega(v_i) - \omega_p(u)$\nllabel{line7}. \If{$\omega(v_i) > 0$\nllabel{line8} } { Mark $v_i$ and set $\omega_p(m_i) \leftarrow \omega_p(m_i) + \omega(v_i)$.\nllabel{line9} } } \For{$i= n~{\bf downto}~1$\nllabel{line10} } { \If{$v_i$ is marked and $m_i$ is not blocked\nllabel{line11} } { Set $D \leftarrow D \cup \{v_i\}$ and block all $u \in N_{G}(v_i)$.\nllabel{line:compMWISend} } } \If{$\sum_{v \in D} \left(M \cdot |N[v]| - w(v)\right)\geq k'$\nllabel{line13} } { \Return{$D$};\nllabel{line14} } \Else{\nllabel{line15} \Return{\sc No};\nllabel{line16} } } \end{algorithm} \begin{theorem}\label{thm:meed} Algorithm~\ref{algo:EDdc} solves the {\sc{Min-WED }} problem for dually chordal graphs in linear time. \end{theorem} \begin{proof} To get Algorithm~\ref{algo:EDdc} we extended the Algorithm~\ref{algo:mwisChordal} by lines~\ref{line:posWeightsStart}--\ref{line:posWeightsEnd} to ensure non-negative weights. Additionally, we added the calculation for~$M$ and~$k'$ (line~\ref{line:compMk}), and defined $\omega(v)$ now as $M \cdot |N[v]| - w(v)$ instead of $|N[v]|$ (line~\ref{line:compOmega}). Therefore, based on Lemma~\ref{lem:wed-mwis}, we first created an instance $(G,w,k)$ for the {\sc{Min-WED }} problem with non-negative weights, and then by Lemma~\ref{lem:wed-mwis-2} an instance $(G,w',k')$ for the {\sc{MWIS }} problem on $G^2$ (with $w'(v)=\omega(v)$). Next, the {\sc{MWIS }} instance is solved in the lines~\ref{line:compMNO}--\ref{line:compMWISend} in linear time. To achieve this, we rewrite Algorithm~\ref{algo:mwisChordal} to work on the dually chordal graph instead of its chordal square. Based on Lemma~\ref{lem:MNO_linear}, line~\ref{line:compMNO} of Algorithm~\ref{algo:EDdc} computes a perfect elimination ordering of $G^2$. Let $v_i$ and $v_j$ be adjacent in $G^2$ and $i <j$. Now Lemma~\ref{lem:ijInE_iff_mjInE2} allows to modify the two loops. For the first loop (line~\ref{line6}), there is an extra vertex weight $\omega_p$. When $v_i$ is marked, instead of decrementing the weights $\omega$ of the neighbors of $v_i$ (and their neighbors), $\omega_p$ of $v_i$'s maximum neighbour~$m_i$ is incremented by $\omega(v_i)$. Now before comparing $\omega(v_j)$ to $0$, $\omega(v_j)$ is decremented by $\omega_p(u)$ for all $u \in N[v_j]$. Because $m_i$ is adjacent to $v_j$ (Lemma~\ref{lem:ijInE_iff_mjInE2}), this ensures that each time the weight~$\omega(v_j)$ is compared to $0$, it has the same value as it would have in Algorithm~\ref{algo:mwisChordal}. For the second loop (line~\ref{line10}) the argumentation works similarly. After selecting a vertex~$v_j$ (i.e., $v_j \in D$), all its neighbors in $G^2$ are blocked. Thus by Lemma~\ref{lem:ijInE_iff_mjInE2}, if a vertex~$v_i$ is adjacent in $G^2$ to a selected vertex, then the maximum neighbor $m_i$ is blocked. It follows that after line~\ref{line:compMWISend}, $D$ is a solution to the instance $(G,w',k')$ for the {\sc{MWIS }} problem on $G^2$ created earlier. Thus, the set~$D$ is a solution for the given {\sc{Min-WED }} instance if and only if $\sum_{v \in D} \big(M \cdot |N[v]| - w(v)\big) \geq k'$ (lines~\ref{line13}--\ref{line16}). \medskip Solving the efficient domination problem (line~\ref{line:compgamma}) and finding a maximum neighborhood ordering (line~\ref{line:compMNO}) can be done in linear time \cite{BCD98,BLR}. Also the overall runtime of lines~\ref{line6}--\ref{line:compMWISend} is bounded by the number of vertices and edges in $G$. Thus, Algorithm~\ref{algo:EDdc} runs in linear time. \end{proof} As explained at the end of the proof of Theorem~\ref{thm:wed-mwis}, the {\sc{Max-WED }} problem can be reduced in time ${\mathcal O}(n+m)$ to the {\sc{Min-WED }} problem by negating the weights. Therefore, the {\sc{Max-WED }} problem can also be solved in linear time on dually chordal graphs. \subsection{AT-free graphs} \label{sec:AT-free} We now consider the class of AT-free graphs. Recall the following result due to Chang {\it et al.}~\cite{CHK03}. \begin{theorem}[\cite{CHK03}]\label{ATGk} Every proper power of an AT-free graph is a cocomparability graph. \end{theorem} Therefore, by Theorem~\ref{thm:wed-mwis}, a polynomial time algorithm for the {\sc{Min-WED }} and {\sc{Max-WED }} problems in AT-free graphs will follow if the {\sc{MWIS }} problem is solvable in polynomial time in the class of cocomparability graphs. The {\sc{MWIS }} problem in the class of cocomparability graphs is equivalent to the maximum weight clique problem in the class of comparability graphs, for which the following result is known. \begin{theorem}[\cite{Gol04,MS99}] The maximum weight clique problem can be solved in linear time on comparability graphs. \end{theorem} The algorithm given by Golumbic~\cite{Gol04} requires a transitive orientation of a comparability graph $G$ as input. Such an orientation can be found in linear time using methods of McConnell and Spinrad~\cite{MS99}. Consequently: \begin{theorem}\label{MWIS:cocomp} The {\sc{MWIS }} problem on instances $(G,w,k)$ such that $G$ is a cocomparability graph can be solved in time ${\mathcal O}(|V(G)|+|E(\overline{G})|)$. \end{theorem} Theorems~\ref{thm:wed-mwis}, \ref{ATGk} and~\ref{MWIS:cocomp} imply the following result. \begin{theorem}\label{thm:AT-free} The {\sc{Min-WED }} and {\sc{Max-WED }} problems are solvable in the class of AT-free graphs in time ${\mathcal O}\big(\min\{nm+n^2,n^\omega\}\big)$. \end{theorem} \begin{proof} Let ${\cal C}$ be the class of AT-free graphs, and let $G\in {\cal C}$ with $n = |V(G)|$. By Theorem~\ref{ATGk}, the square of $G$ is cocomparability. By Theorem~\ref{MWIS:cocomp}, the {\sc{MWIS }} problem is solvable on $G^2$ in time ${\mathcal O}(n^2)$. Thus, Theorem~\ref{thm:wed-mwis} implies that in the class of AT-free graphs, the {\sc{Max-WED }} and {\sc{Min-WED }} problems are solvable in time ${\mathcal O}\big(\min\{nm+n,n^\omega\}+ n^2\big) = {\mathcal O}\big(\min\{nm+n^2,n^\omega\}\big)$. \end{proof} Theorem~\ref{thm:AT-free} generalizes the polynomial-time algorithms for the efficient dominating set problem in the class of AT-free graphs by Brandst\"adt \emph{et al.}~\cite{BLR} and by Broersma \emph{et al.}~\cite{BKM99}. Both papers~\cite{BLR,BKM99} solve the unweighted version of the problem in time ${\mathcal O}(n^4)$, while we give an algorithm of complexity ${\mathcal O}\big(\min\{nm+n^2,n^\omega\}\big)$ to solve the more general, weighted versions of the problem. The polynomial time solvability implied by Theorem~\ref{thm:AT-free} can also be seen as a common extension of the polynomial-time solvability of the {\sc{Min-WED }} problem on interval graphs~\cite{CL94}, cocomparability graphs~\cite{CRC95}, and permutation graphs~\cite{LLT97}, all subclasses of AT-free graphs. \subsection*{Acknowledgements} We are grateful to the two anonymous referees for their helpful comments. M.M.~is supported in part by ``Agencija za raziskovalno dejavnost Republike Slovenije'', research program P$1$--$0285$ and research projects J$1$-$5433$, J$1$-$6720$, and J$1$-$6743$. \bibliographystyle{abbrv}
\section{INTRODUCTION} Ever since the first hard X-ray observations of solar flares \citep{1959JGR....64..697P}, it has been realized that these events are responsible for the acceleration of copious amounts of charged particles, in particular deka-keV electrons. Fifty years of observations have revealed considerable insight into the spectral, temporal, and spatial properties of these accelerated electrons; however, the underlying mechanism responsible for their acceleration remains largely undetermined. A major objective of contemporary high-energy solar physics research is, then, to understand not only the {\it propagation} of accelerated electrons within the source but also the physics of their {\it acceleration}. To do this requires that we obtain information on the hard X-ray emission produced by accelerated electrons with spectral, spatial and temporal resolutions sufficiently precise to probe the emergence of the accelerated electron spectrum from the initial quasi-Maxwellian population. Acquisition of such data was a key element in the design of the {\em RHESSI} instrument \citep{2002SoPh..210....3L}. Proposed acceleration mechanisms include acceleration by large-scale coherent sub-Dreicer electric fields \citep[e.g.,][]{1994ApJ...435..469B} and by supra-Dreicer electric fields in thin reconnecting current sheets \citep[e.g.,][]{1993SoPh..146..127L,1996ApJ...462..997L}. However, these models face serious challenges in terms of the properties of the source (e.g., fine fragmentation, efficient pitch-angle scattering) in order to avoid unacceptably large unidirectional currents \citep{1985ApJ...293..584H,1995ApJ...446..371E}. Further, there is growing observational evidence \citep[e.g.,][]{1980ApJ...239L..85K,2006ApJ...653L.149K} that the overall accelerated electron distribution has an angular distribution that is nearly isotropic. Combined, these theoretical and observational considerations favor a stochastic acceleration model invoking plasma turbulence where particles undergo multiple energetic ``boosts'' by an ensemble of scattering centers \citep[e.g.,][for reviews]{1994ApJS...90..623M, 1997JGR...10214631M, 2012SSRv..173..535P, 2012ApJ...754..103B}. Stochastic acceleration models have been applied to solar flares \citep[e.g.,][]{PhysRev.111.1206,1979AIPC...56..135R,1996ApJ...461..445M,2010ApJ...712L.131P}, and often share the property that the acceleration can be described by a second-order velocity diffusion coefficient $D_{vv}$. We therefore here consider a model that involves acceleration by a stochastic process, modeled through a diffusion term in the Fokker-Planck equation describing the evolution of the electron phase-space distribution function, coupled with particle transport that consists of two components: {\it in situ} Coulomb collisions with the background plasma, and escape associated with the finite length of the acceleration region. In general, the results are characterized by four governing timescales: the acceleration timescale $\tau_{\rm acc}$, the collisional deceleration timescale $\tau_{\rm c}$, the collisional diffusion timescale $\tau_{\rm d}$, and the escape timescale $\tau_{\rm esc}$. In the region in velocity space where escape can be neglected, the electron distribution function is driven toward a steady state corresponding to a balance between diffusive acceleration and collisional energy losses. For a velocity diffusion coefficient $D_{vv} \sim 1/v$, this equilibrium state takes the form of a kappa distribution, which transitions smoothly from a Maxwellian low-energy core to a power-law high-energy tail \citep{0038-5670-9-3-R04,1977ApJ...211..270B, 1985PhRvL..54.2608H,1998GeoRL..25.4099M,2004PhPl...11.1308L}. This result is encouraging, since kappa distributions have been used to characterize particle distribution functions in a variety of space plasma scenarios \citep[e.g.,][]{2009JGRA..11411105L}, including electrons in solar flares \citep{2009A&A...497L..13K,2013ApJ...764....6O} The context and general properties of the acceleration model are detailed in Section~\ref{general}. In Section~\ref{kap}, we consider the steady-state solution for the accelerated electron distribution in the case of an acceleration model characterized by a diffusion coefficient with an inverse dependence on velocity. This takes the form of a kappa distribution, which is characterized by two parameters, one of which is a characteristic velocity scale (e.g., the thermal velocity associated with the Maxwellian core) and the other is the dimensionless ratio of the acceleration time to the collisional deceleration time. The dimensionless parameter, denoted by $\kappa$, is simply related to the power-law spectral index $\delta$ of the electron energy flux, and hence to $\gamma$, the power-law index of the emitted hard X-ray bremsstrahlung spectrum. In Section~\ref{evol} we characterize the time evolution toward the asymptotic kappa distribution as an advancing wavefront in velocity space, and we find that the acceleration of electrons to energy $E$ occurs on a timescale $\tau_{\rm acc} \propto E^{3/2}$. At sufficiently high energies, particle escape associated with the finite length of the acceleration region can modify this asymptotic form, and we consider this effect in Section~\ref{escape-analysis}. In Section~\ref{numerical}, we present numerical solutions of the basic Fokker-Planck equation and we discuss the extent to which the numerical results confirm the analytic results of the previous sections. In Section~\ref{stationary} we analyze a model that is appropriate to acceleration by a coherent large-scale electric field, showing that the presence of efficient pitch-angle scattering can lead to isotropization of the distribution function and hence can produce an effect akin to stochastic acceleration over a wide velocity range. We determine the conditions for the turbulent diffusion coefficient in such a model to take the desired form $D_{vv} \sim 1/v$ and we discuss the constraints that the model imposes on the magnitude of the electric field. In Section~\ref{conclusion} we summarize the results obtained. Overall, our results lead to the characterization, over a wide velocity range, of the evolution of the electron distribution toward its asymptotic form, an analysis pertinent to the study of a variety of stochastic acceleration models that have been associated with hard X-ray production during the impulsive phase of a solar flare. \section{CONTEXT AND GENERAL DESCRIPTION OF THE ACCELERATION MODEL}\label{general} {\em RHESSI} has revealed \citep[e.g.,][]{2008ApJ...673..576X} the presence of coronal hard X-ray flare sources with a background density sufficiently high that the accelerated electrons are collisionally stopped in the corona, rather than streaming through it and impacting on the chromosphere to produce hard X-ray footpoints \citep[cf.][]{2003ApJ...595L.107E}. The spatial distribution of these thick-target coronal hard-X ray sources exhibits a core region where acceleration occurs, surrounded by a halo where escaping high-energy electrons are collisionally stopped \citep{2008ApJ...673..576X}. Since the acceleration and hard X-ray emitting regions are coincident, these flares have opened new horizons for the study of acceleration processes, inasmuch as they permit determination of the length of, and density within, the acceleration region \citep{2008ApJ...673..576X,2011ApJ...730L..22K,2012A&A...543A..53G}, and hence the number of particles available for acceleration and the specific acceleration rate (electrons~s$^{-1}$ per ambient electron), a quantity that measures the efficiency of the acceleration process \citep{2013ApJ...766...28G}. Spectroscopic imaging observations with {\em RHESSI} also suggest the presence of turbulence (due to, e.g., fluctuations in the magnetic field) in these coronal loops \citep{2011ApJ...730L..22K}, resulting in both pitch-angle scattering \citep{2014ApJ...780..176K} and cross-field transport \citep{2011A&A...535A..18B} of high-energy electrons. Our aim is to develop a model for the electron phase-space distribution function in coronal thick-target sources, using a Fokker-Planck equation that includes the combined effects of turbulent acceleration and Coulomb collisions with the dense background plasma: \begin{equation}\label{fund} \frac{\partial f}{\partial t}=\frac{1}{v^{2}} \, \frac{\partial }{\partial v} \, \left \{ v^{2} \left [ \left ( \frac{\Gamma \, v_{\rm te}^{2}}{2v^{3}}+D_{\rm turb}(v) \right ) \, \frac{\partial f}{\partial v}+\frac{\Gamma}{v^{2}} \, f \right ] \right \} \,\,\, . \end{equation} Here $f$ (electrons~cm$^{-3}$~(cm~s$^{-1}$)$^{-3}$) is the phase-space distribution function of electrons, averaged over the acceleration region volume, and we use a simplified form of the collision operator applicable to the solar flare situation \citep[cf.][]{2014ApJ...787...86J}, in which the background electrons are modelled as a heat bath at a fixed temperature $T$(K). In Equation~(\ref{fund}) \begin{equation}\label{vte-def} v_{\rm te}=\sqrt{2k_{B}T/m_{e}} \end{equation} is the thermal speed (with $k_B$ (erg~K$^{-1}$) the Boltzmann constant and $m_e$ (g) the electron mass), \begin{equation}\label{gamma-def} \Gamma = \frac {4\pi e^{4} \ln\Lambda \, n} {m_{e}^{2}} \end{equation} is the collision parameter (with $n$ (cm$^{-3}$) the density of background electrons, $e$ (esu) the electronic charge, and $\ln \Lambda$ the Coulomb logarithm), and $D_{\rm turb}(v) \equiv D_{vv}$ (cm$^2$~s$^{-3}$) is the diffusion coefficient in velocity space associated with an as yet unspecified stochastic acceleration mechanism. There are three characteristic timescales in the stochastic acceleration model represented by Equation~(\ref{fund}), viz. \begin{itemize} \item the {\it acceleration time} $\tau_{\rm acc}$, defined through \begin{equation}\label{tacc-def} \frac{1}{v^{2}} \, \frac{\partial }{\partial v} \, \left \{ v^{2} \left [ D_{\rm turb}(v) \, \frac{\partial f}{\partial v} \right ] \right \} \simeq \frac{f}{\tau_{\rm acc}(v)} \,\,\, ; \,\,\, \tau_{\rm acc}(v)=\frac{v^2}{D_{\rm turb}(v)} \,\, \, ; \end{equation} \item the {\it collisional deceleration/friction time} $\tau_{\rm c}$, defined through \begin{equation}\label{tc-def} \frac{\Gamma}{v^{2}} \, \frac{\partial f}{\partial v} \,\,\, \simeq \frac{f}{\tau_{\rm c}(v)} \,\,\, ; \,\,\, \tau_{\rm c}(v) \simeq \frac{v^3} {\Gamma} \,\,\, ; \qquad {\rm and} \end{equation} \item the {\it collisional diffusion time} $\tau_{\rm d}$, defined through \begin{equation}\label{tdiff-def} \frac{1}{v^{2}} \, \frac{\partial }{\partial v} \, \left \{ v^{2} \left [ \frac{\Gamma v_{te}^{2}}{2v^{3}} \, \frac{\partial f}{\partial v} \right ] \right \} \simeq \frac{f}{\tau_{\rm d}(v)} \, \, \, ; \,\,\, \tau_{\rm d}(v) \simeq \frac{2 v^{5}}{\Gamma v_{te}^{2}} \,\,\, . \end{equation} \end{itemize} Many important properties of the model are more conveniently derived by recasting the Fokker-Planck Equation~(\ref{fund}) in the form studied by \citet{2005PhRvE..72f1106C} and \citet{2010PhyA..389.1021L}: \begin{equation}\label{fp-transformed} \frac{\partial f}{\partial t}=\frac{1}{{v^{2}}} \, \frac{\partial }{\partial v} \, \left [ v^{2} \, D(v) \left ( \frac{\partial f}{\partial v} +f \, U^\prime(v) \right ) \right ] \,\,\, , \end{equation} with \begin{equation}\label{dv} D(v)=\frac{\Gamma \, v_{\rm te}^{2}}{2 \, v^{3}} + D_{\rm turb}(v) \end{equation} and \begin{equation}\label{uprime} U^\prime(v)=\frac{\Gamma}{v^{2}D(v)} = \left ( {v_{\rm te}^2 \over 2v} + { v^2 \, D_{\rm turb}(v) \over \Gamma } \right )^{-1} \,\,\, . \end{equation} Observations of quantities related to $f(v,t)$ are generally averaged over the pertinent instrument time resolution. Specifically, imaging spectroscopy hard X-ray observations from {\em RHESSI} \citep[see][for a review]{2011SSRv..159..301K} are limited to the time it takes to develop a full set of spatial Fourier components of the source; this takes a full spacecraft rotation period of several seconds. Given that the timescales for acceleration, collisional energy loss, and escape are, for typical conditions in loop-top coronal hard X-ray sources \citep{2013ApJ...766...28G}, less than a second, it follows that a quasi-steady-state scenario is of considerable relevance and interest. The stationary ($\partial/\partial t = 0$) solution of the Fokker-Planck equation~(\ref{fp-transformed}) is \begin{equation}\label{ss} f(v)=A \, e^{-U(v)} \,\,\, , \end{equation} where $A$ is a normalization constant. This result is sufficiently general to permit the determination of the steady-state distribution $f(v)$ of energetic electrons in a collisional plasma, given a specific choice of the turbulent velocity-space diffusion coefficient $D_{\rm turb}(v)$ or, equivalently, the function $U(v)$. \section{THE KAPPA DISTRIBUTION AS A STATIONARY SOLUTION}\label{kap} The stationary state (\ref{ss}) has been written in the form of a Gibbs-Boltzmann distribution with the function $U(v)$ playing the role of a potential. It is well known that such distributions globally minimize the Helmoltz free-energy functional $F[f]=E[f]-S[f]$ where $E[f]=\int U f \, d\mathbf{v}$ is the potential energy and $S[f]=-\int f \ln f \, d\mathbf{v}$ is the Boltzmann entropy. This property of the equilibrium state (\ref{ss}) is intimately related to the existence of a variational principle underlying the Fokker-Planck equation (\ref{fp-transformed}), which dictates that the time-dependent solution $f(v,t)$ evolves according to the following constraint \citep{2005PhRvE..72f1106C} on the functional $F[f(v,t)]$: \begin{equation}\label{cst} \dot{F}=-\int \frac{D(v)}{f} \, \left ( \frac{\partial f}{\partial \mathbf{v}}+f \, \frac{\partial U}{\partial \mathbf{v}} \right )^{2}d\mathbf{v} \, \leq \, 0 \,\,\, . \end{equation} Therefore, if $F$ is bounded from below, the distribution function converges toward the stationary state (\ref{ss}) as $t\rightarrow \infty$, corresponding to a statistical equilibrium between diffusive acceleration and collisional drag. Indeed, the electron distribution function will steadily converge toward the stationary state~(\ref{ss}) provided the zero-flux boundary condition $v^{2} \, D(v) (\partial f/\partial v +f \, U^\prime(v) ) \rightarrow 0$ as $v \rightarrow \infty$. When $D_{\rm turb}=0$, Equation~(\ref{uprime}) shows that the potential $U(v)$ is quadratic in $v$, so that, by Equation~(\ref{ss}), a steady-state Maxwellian distribution is obtained. In this case the Fokker-Planck equation describes collisional relaxation toward thermal equilibrium at temperature $T$. However, Equation~(\ref{uprime}) also shows that a steady-state Maxwellian distribution of electrons can be achieved in the presence of a finite level of turbulence $D_{\rm turb}\neq 0$ provided $D_{\rm turb} \sim 1/v^3$, in which case the acceleration time given by Equation~(\ref{tacc-def}) obeys a velocity dependence\footnote{acceleration times $\tau_{\rm acc} \sim v^{5}$, corresponding to $D_{\rm turb}\sim v^{-3}$, are produced by Gaussian isotropic spectra of electrostatic fluctuations, not necessarily thermal, in the plasma \citep[see][]{1992PhRvL..69.1831R}} identical to that of collisional diffusion $\tau_{acc}(v)\sim \tau_{d}(v) \sim v^{5}$. Therefore, the presence of a Maxwellian distribution of plasma electrons is \emph{not} synonymous with a state of thermal equilibrium. This fact may complicate the interpretation of spectroscopic data, inasmuch as the temperature inferred from the shape of the electron distribution function may also include a turbulent broadening component \citep[e.g.,][]{1986ApJ...301..975A}. The distribution function of deka-keV electrons in solar flares is generally well described by a Maxwellian core with a power-law high-energy tail \citep[see, e.g.,][]{2003ApJ...595L..97H}. In an attempt to account for this behavior, let us consider a turbulent diffusion coefficient of the form \begin{equation}\label{dturb-result} D_{\rm turb} (v) = {D_{0} \over v} \,\,\ , \end{equation} from which it follows (Equation~(\ref{tacc-def})) that the acceleration time, defined as $\tau_{\rm acc} (v) \equiv {v^2 / D(v)}$, is given by \begin{equation}\label{tau-acc-v3d0} \tau_{\rm acc} (v) = {v^3 \over D_0} \,\,\, . \end{equation} For this case, the acceleration time $\tau_{\rm acc}$ and the collisional deceleration time $\tau_{\rm c}$ have the same velocity dependence, $\tau_{\rm acc}(v) \propto \tau_{\rm c}(v) \propto v^3$. Therefore we can define the dimensionless constant \begin{equation}\label{kappadef} \kappa = {\tau_{\rm acc}(v) \over 2 \, \tau_{\rm c}(v)} = \frac{\Gamma}{2 D_{0}} \,\,\, , \end{equation} (the reason for the factor 2 will be evident shortly). With this identification, Equation~(\ref{uprime}) becomes \begin{equation}\label{uprime-1-over-v} U'(v) = {2 v \over v_{\rm te}^2} \left ( 1 + {v^2 \over \kappa \, v_{\rm te}^2 } \right )^{-1} \,\,\, , \end{equation} with the following solution for the potential $U(v)$: \begin{equation}\label{u-solution} U(v) = \kappa \, \ln \left ( 1 + {v^2 \over \kappa \, v_{\rm te}^2} \right ) \,\,\, . \end{equation} Thus, by Equation~(\ref{ss}), the (normalized) stationary solution is the well-known \citep[see, e.g.,][]{1968JGR....73.2839V,2009A&A...497L..13K,2009JGRA..11411105L,2013ApJ...764....6O} kappa distribution \begin{equation}\label{kappa-dist} f_\kappa(v)= \frac{n_{\kappa}}{\pi^{3/2} \, v_{\rm te}^{3} \, \kappa^{3/2}} \, \frac{\Gamma(\kappa)}{\Gamma \left (\kappa-\frac{3}{2} \right ) } \, \left ( 1+\frac{v^{2}}{\kappa \, v_{\rm te}^{2}} \right )^{-\kappa} \,\,\, , \end{equation} where $n_\kappa = \int f_{\kappa}(v) \, d^{3}v$ is the number density associated with the accelerated electron distribution. Equation~(\ref{kappa-dist}) defines a {\it kappa distribution of the first kind}, in the terminology of \citet[][their Equation~(9)]{2009JGRA..11411105L}. We note that other authors \citep[e.g.,][]{2009A&A...497L..13K,2013ApJ...764....6O} have used a kappa distribution of the {\it second} kind \citep[again in the terminology of][their Equation~(10)]{2009JGRA..11411105L}: \begin{equation}\label{kappa-dist-second-kind} f_{\tilde \kappa}(v)= \frac{n_{\tilde \kappa}}{\pi^{3/2} \, \theta^{3} \, {\tilde \kappa}^{3/2}} \, \frac{\Gamma({\tilde \kappa}+1)}{\Gamma \left ({\tilde \kappa}-\frac{1}{2} \right ) } \, \left ( 1+\frac{v^{2}}{{\tilde \kappa} \, \theta^2} \right )^{-({\tilde \kappa}+1)} \end{equation} to describe the electron distribution function in solar flares. Such authors have also used the concept of {\it kinetic temperature} $T_K$, defined such that the average energy of the electrons in the kappa distribution (\ref{kappa-dist-second-kind}) is ${\overline E} = (3/2) \, k_{B} \, T_K$ \citep[see, e.g.,][]{2013ApJ...764....6O}. It follows that \begin{equation}\label{kinetic_temperature} k_{B}T_K={1 \over 2} \, m_e \, \theta^2 \left [{{\tilde \kappa} \over ({\tilde \kappa} - 3/2)}\right ] \end{equation} and it should be noted that the kinetic temperature $T_K$ is not to be confused with $T$ (Equation~(\ref{vte-def})), the temperature of the background Maxwellian with which the accelerated electrons interact. It is important to note that Equations~(\ref{kappa-dist}) and~(\ref{kappa-dist-second-kind}) refer to an identical family of two-parameter distributions; only the parametric labelling of the mathematical form is different in the two descriptions. Indeed, as noted by \citet{2009JGRA..11411105L}, the changes of variable \begin{equation}\label{change_of_paramters} {\tilde \kappa} = \kappa - 1 ; \qquad \theta = \sqrt{\kappa \over \kappa - 1} \,\, v_{\rm te} \end{equation} transform Equation~(\ref{kappa-dist}) into Equation~(\ref{kappa-dist-second-kind}) {\it exactly}. \citet{2009JGRA..11411105L} note that ``the first kind of kappa distribution is less widely used than the second kind.'' However, in our ``first kind'' parametrization~(\ref{kappa-dist}), the quantity $\kappa$ has an immediate physical significance, namely the dimensionless ratio (Equation~(\ref{kappadef})) of two physical quantities: the stochastic acceleration time $\tau_{\rm acc}$ (Equation~(\ref{tacc-def})) and the collisional deceleration time $\tau_{\rm c}$ (Equation~(\ref{tc-def})) or, equivalently, the collisional parameter $\Gamma$ (Equation~(\ref{gamma-def})) and the diffusion parameter $D_0$ (Equation~(\ref{dturb-result})). We therefore submit that the form~(\ref{kappa-dist}) is a more natural choice of kappa distribution parametrization. Examples of kappa distributions (\ref{kappa-dist}), for various values of the parameter $\kappa$, are shown in Figure~\ref{kappas}. For high values of $\kappa$, the identity \begin{equation} e^{-x} = \lim_{\kappa \rightarrow \infty} \left ( 1 + {x \over \kappa} \right )^{-\kappa} \end{equation} shows that $f_\kappa(v)$ (Equation~(\ref{kappa-dist})) approaches the Maxwellian form \begin{equation}\label{maxwellian} f_\kappa(v) \sim \exp \left ( - {v^2 \over v_{\rm te}^2} \right ) \,\,\, . \end{equation} At low velocities $v \ll \sqrt{\kappa} \, v_{\rm te}$, the collisional diffusion term $f/\tau_{\rm d} \sim v^{-5}$ is dominant over the turbulent term $f/\tau_{\rm acc} \sim v^{-3}$, and so the distribution relaxes through collisional diffusion to a Maxwellian form. Equation~(\ref{kappa-dist}) confirms that in this regime the kappa distribution approaches the form \begin{equation}\label{low-velocity-limit} f \sim \left ( 1 - {v^2 \over v_{\rm te}^2} \right ) \quad {\rm as} \quad v \rightarrow 0 \,\,\, , \end{equation} which is the same as the low-velocity limit of the Maxwellian distribution~(\ref{maxwellian}). On the other hand, in the high-velocity limit $v \gg \sqrt{\kappa} \, v_{\rm te}$, the collisional diffusion timescale $\tau_{\rm d} \sim v^5$ (Equation~(\ref{tdiff-def})) is much longer than either the acceleration time $\tau_{\rm acc}$ (Equation~(\ref{tacc-def})) or the collisional deceleration timescale $\tau_{\rm c}$ (Equation~(\ref{tc-def})), both of which vary with velocity like $v^3$. Thus in this regime the (temperature-dependent) collisional diffusion term is unimportant. Further, since both $\tau_{\rm acc}$ and $\tau_{\rm c}$ have the same velocity dependence ($\sim v^3$), there is no characteristic velocity scale in this domain. Indeed, Equation~(\ref{kappa-dist}) confirms that the stationary distribution approaches a (scale-independent) power-law form: \begin{equation}\label{kappa-power} f_\kappa \rightarrow v^{-2 \kappa} \quad {\rm as} \quad v \rightarrow \infty \,\,\, . \end{equation} Overall, then, the use of the turbulent diffusion coefficient of the form~(\ref{dturb-result}) leads to an accelerated electron distribution that has the form of a kappa distribution~(\ref{kappa-dist}). Such a distribution, as intended, accounts for the observed \citep[e.g.,][]{2003ApJ...595L..97H} blend of a Maxwellian core at low energies with a power law at higher energies. No artificial ``low-energy cutoff'' to the high-energy part of the distribution need be invoked; the electron distribution transitions smoothly from a ``non-thermal'' shape at high energies to a thermal (Maxwellian) form at low energies. Recalling our remarks near the beginning of this Section on the possibility of a stationary Maxwellian form for $f(v)$ even in the presence of finite non-thermal turbulence, it should be noted that {\it any} scenario in which the turbulent acceleration time $\tau_{\rm acc}$ varies between $\tau_{\rm acc}\sim v^{5}$ at low velocities and $\tau_{\rm acc}\sim v^{3}$ at larger velocities will produce a kappa distribution. Thus, we again note that the Maxwellian core of the kappa distribution is not necessarily associated with a collisionally-dominated thermal equilibrium state. Now, the electron phase-space distribution function $f(v)$ is related to the mean electron flux ${\overline F}(E)$ (electrons~cm$^{-2}$~s$^{-1}$ per unit energy) through the relation $v f(v) \, d^3v = {\overline F}(E) \ dE$. Using the elementary relation $E= m_e v^2/2$, it follows that $v \, d^3 v \sim v^3 \, dv \sim E \, dE$ and hence that $f(v) \sim {\overline F(E)}/E$. Thus $\kappa$ is simply related to the power-law spectral index for the mean electron flux: ${\overline F}(E) \sim E^{-\delta}$, with $\delta = \kappa-1$. \begin{figure}[htpb] \centering \includegraphics[width=10cm]{f1.eps} \caption{The stationary solution kappa distribution, $f_{\kappa}$, for different values of $\kappa$, all normalized to a density $n_\kappa=1$. {\it Solid blue line:} $\kappa = 1.6$, {\it dotted orange line:} $\kappa = 3$, {\it dashed green line:} $\kappa = 5$, {\it dot-dashed red line:} $\kappa = 10$, {\it dot-dot-dot-dashed purple line:} $\kappa = 30$. For small values of $\kappa$ the distribution function has a Maxwellian core and a non-thermal power-law tail, while for large values of $\kappa$, the distribution is almost indistinguishable from a Maxwellian.} \label{kappas} \end{figure} Observations show that the hard X-ray spectrum above $\sim$(15-20)~keV is indeed approximately power-law in form: $I(\epsilon) \sim \epsilon^{-\gamma}$, with a typical value $\gamma \simeq$~5. The bremsstrahlung hard X-ray spectrum $I(\epsilon)$ (photons~cm$^{-2}$~s$^{-1}$~keV$^{-1}$ at the Earth) is related to the emitting mean electron flux spectrum ${\overline F}(E)$ (electrons~cm$^{-2}$~s$^{-1}$~keV$^{-1}$) \citep{2003ApJ...595L.115B} by \begin{equation}\label{xray-electron} I(\epsilon) = {n \, V \over 4 \pi R^2} \int_\epsilon^\infty {\overline F}(E) \, \sigma(\epsilon, E) \, dE \,\,\, , \end{equation} where $V$ is the source volume, $R$ = 1 AU, and $\sigma(\epsilon,E)$ is the bremsstrahlung cross-section (cm$^2$~keV$^{-1}$), differential in photon energy $\epsilon$. For the simple non-relativistic Kramers cross-section \begin{equation}\label{kramers} \sigma(\epsilon, E) \sim {1 \over \epsilon E} \,\,\, , \end{equation} a hard X-ray spectrum $I(\epsilon) \sim \epsilon^{-\gamma}$ thus implies a mean electron flux spectrum ${\overline F}(E) \sim E^{-\delta}$, with $\delta = \gamma -1$; this relation also holds for more complex forms of $\sigma(\epsilon, E)$, such as the Bethe-Heitler cross-section \citep[see][]{1971SoPh...18..489B}. As discussed above, for the model considered here, the electron flux at high energies approximates a power-law with $\delta = \kappa-1$; thus the hard X-ray spectral index $\gamma$ and the electron distribution parameter $\kappa$ are equal: \begin{equation}\label{kappa-gamma} \kappa = \gamma \,\,\, , \end{equation} and so a typical value of $\kappa \simeq 5$. Further, to obtain such a value of $\kappa$, Equation~(\ref{kappadef}) shows that the acceleration time \begin{equation}\label{ratio-10} {\tau_{\rm acc} } \simeq 10 \, \tau_{\rm c} \,\,\, , \end{equation} i.e., about an order of magnitude larger than the collisional friction/deceleration time. Since the power-law index $\kappa$ in this acceleration model is proportional to the acceleration time (Equation~(\ref{kappadef})), it follows that temporal hardening (softening) of the photon spectrum can be produced by a decrease (increase) of the acceleration time, resulting from a variation of the turbulent diffusion coefficient $D_0$ on a time-scale much longer than the overall relaxation time toward the steady state. The same argument was advanced by \citet{1977ApJ...211..270B} for interpreting spectral index variations of the photon spectrum during solar flares, including the commonly observed soft-hard-soft behavior. \section{EVOLUTION TOWARD THE STATIONARY DISTRIBUTION}\label{evol} We now consider in more detail the relaxation of the electron distribution function toward the stationary solution~(\ref{kappa-dist}), and in particular the formation of accelerated high-energy tails during such a process. This can be studied by introducing the function \begin{equation}\label{udef} u(v,t) \equiv \frac{f(v,t)}{f_\kappa (v)} \,\,\, . \end{equation} Substituting $f(v,t) = f_{\kappa}(v) \, u(v,t)$ into Equation~(\ref{fp-transformed}) and using the fact that $\partial f_\kappa /\partial t = 0$, we obtain an equation governing the evolution of the dimensionless quantity $u(v,t)$: \begin{equation}\label{u-evol} \frac{\partial u}{\partial t}=\frac{1}{v^{2}} \, \frac{\partial}{\partial v} \left ( v^{2} \, D(v) \, \frac{\partial u}{\partial v} \right ) -D(v) \, U^\prime(v) \, \frac{\partial u}{\partial v} \,\,\, . \end{equation} We now introduce the velocity-space variable $\eta$ through the transformation \begin{equation}\label{yvtrans} d\eta = \frac{dv}{\sqrt{D(v)}} \end{equation} and thus find that Equation~(\ref{u-evol}) can be written in the form of an advection-diffusion equation in velocity space: \begin{equation}\label{fp-diffusion} \frac{\partial u}{\partial t} + V(v) \, \frac{\partial u}{\partial \eta}=\frac{\partial^{2}u}{\partial \eta^{2}} \,\,\, . \end{equation} Here the advection speed (in velocity space) is given by \begin{equation}\label{advection-speed} V(v)=\sqrt{D(v)} \left [ U^\prime(v)-\frac{2}{v}-\frac{1}{2} \, {d \ln D(v) \over dv} \right ] \,\,\, . \end{equation} Because of the advection-diffusion structure of the Equation~(\ref{fp-diffusion}) that governs the relaxation toward the kappa distribution, the acceleration process is characterized by the successive energization of particles of higher and higher energy. The process may thus be described as a velocity-space ``front'' moving in the direction of increasing velocity \citep{1957PhRv..107..350M, 1964pkt..book.....M, 2005PhRvE..72f1106C}. The position $v_{f}(t)$ of this velocity-space front may be estimated by neglecting the diffusion term in Equation~(\ref{fp-diffusion}), so that \begin{equation}\label{partialu-partialt} \frac{\partial u}{\partial t} + V(v_{f}) \, \frac{\partial u}{\partial \eta} = 0 \,\,\, . \end{equation} The location of the velocity-space ``front'' may be identified with a fixed value of $u(v,t)=f(v,t)/f_\kappa$ (see Section~\ref{numerical}, where we set $u=0.5$). Thus, setting the total derivative \begin{equation}\label{du-total} {du \over dt} \equiv \frac{\partial u}{\partial t} + {d \eta \over dt} \, \frac{\partial u}{\partial \eta} = 0 \end{equation} allows us to write \begin{equation}\label{vvf} V(v_{f}) = \frac{d\eta}{dt} = \frac{1}{\sqrt{D(v_{f})}} \, \frac{dv_{f}}{dt} \end{equation} and hence \begin{equation}\label{dvfdt} \frac{dv_{f}}{dt} = \sqrt{D(v_{f})} \, V(v_{f}) = D(v_{f}) \, \left [ U^\prime(v_{f})-\frac{2}{v_{\rm f}} - \frac{1}{2 \, v_{f}} \, {d \ln D(v_{f}) \over d \ln v_{f}} \right ] \,\,\, . \end{equation} In the high-velocity domain, $D(v) \simeq D_{\rm turb}(v) = D_0/v$ (Equations~(\ref{dv}) and~(\ref{dturb-result})) so that $d \ln D(v_f)/d \ln v_f = -1$. Also, from Equation~(\ref{uprime-1-over-v}), in this regime $U^\prime(v_f) \simeq 2 \kappa /v_f$, so that Equation~(\ref{dvfdt}) reduces to \begin{equation}\label{dvfdt-kappa} \frac{dv_f}{dt} = {D(v_f) \over v_f} \, \left ( 2 \, \kappa - \frac{3}{2} \right ) = \Gamma \left ( 1-\frac{3}{4\kappa} \right ) \, \frac{1}{v_f^{2}} \,\,\, , \end{equation} where we have used Equations~(\ref{dturb-result}) and~(\ref{kappadef}). This has solution \begin{equation}\label{vf} v_{f}(t) = \left ( 1 - \frac{3}{4\kappa} \right )^{1/3} \, (3 \, \Gamma t)^{1/3} \, \simeq \, v_{\rm te} \left ( {t \over \tau} \right )^{1/3} \,\,\, , \end{equation} where $\tau$ is the characteristic collision time for a thermal electron: \begin{equation}\label{tau} \tau = {v_{\rm te}^3 \over 3 \, \Gamma} = {(2kT)^{3/2} \, m_e^{1/2} \over 12 \pi n e^4 \ln \Lambda} \simeq \, 4 \times 10^{-3} \, {T^{3/2} \over n} \,\,\, . \end{equation} Substituting typical numerical values for the flaring corona in a dense looptop source, viz. $T= 2 \times 10^7$~K, $n=10^{11}$~cm$^{-3}$, we obtain $\tau \simeq 3 \, {\rm ms}$. A hard-X-ray-producing electron has a typical energy $\sim$30~keV, about 15 times the thermal energy. From Equation~(\ref{vf}) we see that \begin{equation} t = \tau \, \left ( {v_f \over v_{\rm te}} \right )^3 \,\,\, , \end{equation} and hence the time to produce an electron of this energy is $\sim (15)^{3/2} \, \tau \simeq 0.2$~s, comparable to the observed rise and decay times of the hard X-ray flux at such energies. This therefore raises the question of whether electrons can be confined in the acceleration region for a time sufficiently long for the ensemble to attain the asymptotic kappa distribution form~(\ref{kappa-dist}). We explore the consequences of this situation more fully in the following section. \section{SPATIAL TRANSPORT AND ESCAPE}\label{escape-analysis} In the acceleration model considered above, it is implicitly assumed that the electron distribution function is maintained close to isotropy as a result of efficient angular scattering in the acceleration region. This implies that the transport of electrons in this region is characterized by a spatial diffusion over length-scales much larger than their mean free-path $\lambda(v)$. Thus, after averaging over the fast pitch-angle scattering time-scale $\tau_{\rm pa}(v)\sim \lambda(v)/v$ responsible for isotropization of the distribution function, the pitch-angle-dependent streaming transport of electrons parallel to the background magnetic field, described by the relation \begin{equation} \dot{z}=\mu \, v \end{equation} (where $z$ is the coordinate along a direction parallel to the guiding magnetic field and $\mu$ is the cosine of the angle between the velocity and magnetic field vectors), assumes the diffusive form \begin{equation}\label{kappa-par} \mu \, v \, \frac{\partial f}{\partial z} \rightarrow \frac{\partial}{\partial z} \, \left [ K_{\parallel} \, \frac{\partial f}{\partial z} \right ] \,\,\, . \end{equation} where the corresponding spatial diffusion coefficient is given by \begin{equation} K_\parallel = \frac{\lambda (v) \, v}{3} \,\,\, . \end{equation} In this strong scattering limit \citep[e.g.,][]{2012SSRv..173..535P}, the Fokker-Planck equation, including the spatial transport term, takes the form \begin{equation}\label{fp-kappa-parallel} \frac{\partial f}{\partial t} + \frac{\partial}{\partial z} \, \left [ {\lambda(v) \, v \over 3} \, \frac{\partial f}{\partial z} \right ] =\frac{1}{v^{2}} \, \frac{\partial }{\partial v} \, \left \{ v^{2} \left [ \left ( \frac{\Gamma \, v_{\rm te}^{2}}{2v^{3}}+D_{\rm turb}(v) \right ) \, \frac{\partial f}{\partial v}+\frac{\Gamma}{v^{2}} \, f \right ] \right \} \,\,\, . \end{equation} Now, representing $\partial/\partial z$ as $1/L$, thus defining the ``length'' $L$ of the acceleration region, Equation~(\ref{fp-kappa-parallel}) can be written \begin{equation}\label{escape} \frac{\partial f}{\partial t}=\frac{1}{v^{2}} \, \frac{\partial }{\partial v} \, \left \{ v^{2} \left [ \left ( \frac{\Gamma \, v_{\rm te}^{2}}{2v^{3}}+D_{\rm turb}(v) \right ) \, \frac{\partial f}{\partial v}+\frac{\Gamma}{v^{2}} \, f \right ] \right \}-\frac{f}{\tau_{\rm esc}(v)} \,\,\, , \end{equation} where the escape time-scale \begin{equation}\label{tau-esc-v} \tau_{\rm esc}(v) = {3 L^2 \over \lambda(v) \, v} = \left ( {3L \over \lambda(v)} \right ) \, \left ( \frac{L}{v} \right ) \,\,\, . \end{equation} In this leaky-box approximation, intended to represent the effect of spatial transport out of the acceleration region, the role of the escape term is to deplete the number of electrons from the acceleration region over a transport time scale $\tau_{\rm esc}(v)$. We notice that this diffusive escape time becomes of the order of the free-streaming escape time $L/v$ only when the mean free path $\lambda$ and the acceleration region length $L$ are comparable. We also note that in the absence of an additional source of particles maintaining a steady state, the number of electrons will decrease with time as a result of the escape term. \begin{figure}[htpb] \centering \includegraphics[width=10cm]{f2.eps} \caption{Characteristic timescales of the system. The solid blue line represents the collisional diffusion timescale $\tau_{\rm d} \propto v^5$ (Equation~(\ref{tdiff-def})), the purple triple-dot-dash line the acceleration timescale $\tau_{\rm acc} \propto v^3$ (Equations~(\ref{tacc-def}) and~(\ref{dturb-result})), the green dashed line the collisional deceleration timescale $\tau_{\rm c} \propto v^3$ (Equation~(\ref{tc-def})), and the red dot-dashed lines the escape time $\tau_{\rm esc} \propto v^{-1}$ (Equation~(\ref{tau-esc-v})) for (from bottom to top) $\lambda/L = 0.2$, $0.01$ and $0.001$. } \label{char-times-fig} \end{figure} In equation (\ref{escape}), there are now four terms (acceleration, collisional deceleration, collisional diffusion, and escape), each with their associated characteristic timescale. The relative importance of these terms is summarized on Figure~2. Ignoring for the moment the (red dash-dot) lines representing the escape time $\tau_{\rm esc}(v)$ ($\sim v^{-1}$ for a velocity-independent mean free path $\lambda$), we can see the two regimes that define the boundaries of the kappa distribution. At low velocities the collisional diffusion time, $\tau_{\rm d}(v) \sim v^5$ (blue line), is shorter than, and hence dominant over, the acceleration timescale $\tau_{\rm acc}(v) \sim v^3$ (purple triple-dot-dash line); this creates a collisionally-dominated Maxwellian core. At higher velocities, the physics is dominated by the acceleration and collisional friction timescales $\tau_{\rm acc}(v)$ and $\tau_{\rm c}(v)$, which have the same velocity dependence $\sim v^3$. The resulting absence of characteristic velocity in this regime yields a power-law spectrum with index $\kappa=\tau_{\rm acc}/2\tau_{\rm c}$. Since the mean free path $\lambda(v)$ can generally be expected to be constant or increase with $v$, the escape time scale $\tau_{\rm esc}(v)$ will also generally be a decreasing function of $v$. Thus, at sufficiently large velocities, $\tau_{\rm esc}$ will eventually become smaller than all of $\tau_{\rm d}(v)$, $\tau_{\rm acc}(v)$, and $\tau_{\rm c}(v)$ (all of which are increasing functions of $v$). Hence we define the escape velocity $v_{\rm esc}$ as the critical velocity where escape starts to be the leading effect, found by equating $\tau_{\rm esc}(v)$ (Equation~(\ref{tau-esc-v})) and $\tau_{\rm c}(v)$ (Equation~(\ref{tc-def})): \begin{equation} {3 L^2 \over \lambda(v_{\rm esc}) \, v_{\rm esc} } = \frac{v_{\rm esc}^3} {\Gamma} \,\,\, . \end{equation} For $\lambda = \lambda_0 (v/v_0)^{-\alpha}$ (see discussion in Section~\ref{stationary}), the explicit solution is \begin{equation} v_{\rm esc} = \left ( {3 L^2 \, \Gamma \over \lambda_0 \, v_0^\alpha} \right )^{1 \over 4 - \alpha} \,\,\, . \end{equation} The three red dot-dashed lines in Figure~\ref{char-times-fig} show $\tau_{\rm esc}(v)$ for three different values of $\lambda/L$ (where $\lambda$ is assumed to be independent of velocity, i.e. $\alpha = 0$). As the mean free path decreases the escape time becomes longer (Equation~(\ref{tau-esc-v})) and thus the intersection with $\tau_{\rm c}(v)$ occurs at a higher velocity. For example, for $\lambda/L = 0.2$, $\tau_{\rm esc}(v)$ intercepts $\tau_{\rm c}(v)$ before the acceleration timescale has become shorter than the collisional diffusion timescale, $\tau_{\rm d}(v)$, so in this case we don't expect a kappa distribution to form. On the other hand, looking at the $\lambda/L = 0.0001$ line we see that the escape time intercepts at much larger velocities, so that a kappa distribution power law tail will form for $v \ll v_{\rm esc}$, the shape of the distribution function becoming substantially different from a kappa distribution only at large velocities $v\gapprox v_{\rm esc}$. We may solve the ``leaky box'' Fokker-Planck equation (\ref{escape}) using an approximation based on an analogy to the pitch-angle loss-cone in a magnetic trap where in the loss-cone situation, there is a critical pitch angle below which electrons escape and above which they remain fully trapped. By analogy, the ``escape velocity'' $v_{\rm esc}$ is the velocity below which electrons are considered to remain in the acceleration region and above which they are considered to freely escape\footnote{this approximation is also used to model the escape of stars from gravitational clusters \citep{1943ApJ....97..263C,1958ApJ...127..544S}}. The Fokker-Planck equation may therefore be written without an explicit escape term: \begin{equation}\label{escc} \frac{\partial f}{\partial t}=\frac{1}{v^{2}} \, \frac{\partial }{\partial v} \, \left \{ v^{2} \left [ \left ( \frac{\Gamma \, v_{te}^{2}}{2v^{3}}+D_{\rm turb}(v) \right ) \, \frac{\partial f}{\partial v}+\frac{\Gamma}{v^{2}} \, f \right ] \right \} \,\,\, , \end{equation} together with the absorbing boundary condition \begin{equation}\label{boundary} f(v_{\rm esc},t)=0 \end{equation} replacing the escape term. We treat the problem in the limit of a small escape rate when the acceleration region can effectively be considered a collisional thick target. Thus the time-dependent solution can be found by perturbation analysis. We start with the Fokker-Planck equation in the form~(\ref{fp-transformed}), repeated here: \begin{equation}\label{fokker-repeated} \frac{\partial f}{\partial t} = \frac{1}{{v^{2}}} \, \frac{\partial}{\partial v} \, \left [ v^{2} \, D(v) \left ( \frac{\partial f}{\partial v} + f \, U^\prime(v) \right ) \right ] \,\,\, , \end{equation} which is to be solved subject to the boundary condition~(\ref{boundary}). We posit a solution of the form \begin{equation}\label{eigenmode-sol} f(v,t)=A \, e^{\nu t} \, g(v) \,\,\, , \end{equation} leading to \begin{equation}\label{eigenvalue-equation} \nu \, g(v) =\frac{1}{{v^{2}}} \, \frac{\partial }{\partial v} \, \left [ v^{2}D(v) \left ( \frac{dg(v)}{dv} +g(v) \, \frac{dU}{dv} \right ) \right ] \,\,\, . \end{equation} This has a first integral \begin{equation}\label{first-integral} \frac{dg(v)}{dv} + g(v) \, \frac{dU(v)}{dv} = \frac{\nu}{v^{2} D(v)} \, \int_{0}^{v}dw \, w^{2} \, g(w) \,\,\, . \end{equation} We next write the solution as an expansion in the decay rate $\nu$ \citep[cf.][]{1965AJ.....70..376K,2010PhyA..389.1021L}: \begin{equation}\label{expamsion} g(v)=g_{0}(v)+\nu \, g_{1}(v) + \ldots \,\,\, . \end{equation} The zero-order equation is \begin{equation}\label{zero-order} \frac{dg_{0}(v)}{dv}+g_{0}(v) \, \frac{dU(v)}{dv} = 0 \,\,\, , \end{equation} with the expected solution (see Equation~(\ref{ss})) \begin{equation}\label{zero-order-soluton} g_{0}(v) = A \, e^{-U(v)} \,\,\, , \end{equation} where $A$ is a normalization factor. The first-order equation is \begin{equation}\label{first-order} \frac{dg_{1}(v)}{dv} + g_{1}(v) \, \frac{dU(v)}{dv} = \frac{1}{v^{2}D(v)}\int_{0}^{v} dw \, w^{2} \, g_{0}(w) = \frac{A}{v^{2}D(v)}\int_{0}^{v} dw \, w^{2} \, e^{-U(w)} \,\,\, . \end{equation} Using an integrating factor $e^{U(v)}$, we derive the solution \begin{equation}\label{first-order-solution} g_{1}(v) = A \, e^{-U(v)} \, \chi(v) \,\,\, , \end{equation} where $\chi(v)$ is the function defined by \begin{equation}\label{chi-def} \chi^\prime (v) = \frac{e^{U(v)}}{v^{2}D(v)} \int_{0}^{v} dw \, w^{2} \, e^{-U(w)} \,\,\, . \end{equation} Using Equations~(\ref{expamsion}), (\ref{zero-order-soluton}), and~(\ref{first-order-solution}), the distribution function is, to first order in $\nu$, given by \begin{equation}\label{time-dept-soln} f(v,t) = A \, e^{-U(v)} \, e^{\nu t} \, [ \, 1 + \nu \, \chi(v) \, ] \,\,\, . \end{equation} Now introducing the boundary condition $f(v_{\rm esc},t)=0$ (Equation~(\ref{boundary})), we obtain the identification \begin{equation}\label{nu-time-soln} \nu = - \, \frac{1}{\chi(v_{\rm esc})} \,\,\, , \end{equation} which is the sought-after escape rate in the limit of large escape velocity. In the case where $D_{\rm turb}(v)=D_{0}/v$ (Equation~(\ref{dturb-result})), we recall that (cf. Equation~(\ref{dv})) \begin{equation}\label{dv-time-soln} D(v)=\frac{\Gamma \, v_{\rm te}^{2}}{2 \, v^{3}}+\frac{D_{0}}{v} \,\,\, , \end{equation} and that (Equation~(\ref{u-solution})) \begin{equation}\label{uv-time-soln} U(v)=\kappa \ln \left ( 1 + { v^2 \over \kappa v_{\rm te}^2} \right ) \,\,\, . \end{equation} Substituting results~(\ref{nu-time-soln}) and~(\ref{uv-time-soln}) in Equation~(\ref{time-dept-soln}) gives the (normalized; see Equation~(\ref{kappa-dist})) time-dependent solution of the leaky-box acceleration model: \begin{equation}\label{leaky-box-soln} f(v,t)= \frac{n \, e^{-t/\chi(v_{\rm esc})}}{\pi^{3/2} \, v_{te}^{3} \, \kappa^{3/2}} \, \frac{\Gamma(\kappa)}{\Gamma \left (\kappa-\frac{3}{2} \right ) } \, \left ( 1+\frac{v^{2}}{\kappa \, v_{\rm te}^{2}} \right )^{-\kappa} \left [ 1-\frac{\chi(v)}{\chi(v_{\rm esc})} \right ] \,\,\, . \end{equation} The last factor in brackets describes the deviation from the kappa distribution and \begin{equation} n(t)=n \, \exp[{-t/\chi(v_{\rm esc})}] \end{equation} describes the decreasing overall number of particles in the box with time, which are both a consequence of escape of particles out of the acceleration region. These functions depend on the function $\chi(v)$, which is determined through Equations~(\ref{chi-def}), (\ref{dv-time-soln}), and~(\ref{uv-time-soln}): \begin{equation}\label{chi-expression} \chi^\prime(v) = {v \over \kappa \, D_0 \, v_{\rm te}^2} \left ( 1 + \frac{v^{2}}{\kappa \, v_{\rm te}^{2}} \right )^{\kappa-1} \, \int_{0}^{v} dw \, w^{2} \left ( 1+\frac{w^{2}}{\kappa \, v_{\rm te}^{2}} \right )^{-\kappa} \,\,\, . \end{equation} As a reminder, the above solution is valid in the limit where the acceleration region behaves essentially as a thick target. A stationary solution of a similar leaky-box Fokker-Planck equation, without the collisional diffusion term but with a source of particles, was also obtained by \citet{1977ApJ...211..270B}. \section{NUMERICAL SOLUTIONS}\label{numerical} \begin{figure}[htpb] \centering \includegraphics[width=10cm]{f3.eps} \caption{Temporal evolution of electron distribution function $f(v,t)$, for $\kappa \, (\equiv \Gamma/2D_0) = 5$. The solid blue line shows the initial Maxwellian and then, from left to right, $f(v,t)$ at $t/\tau_c = 1.0$ (orange dotted line), $t/\tau_c =10$ (green dashed line), $t/\tau_c=100$ (red dot-dashed line), and $t/\tau_c=1000$ (purple dot-dot-dot-dashed line). }\label{dist-evol-fig} \end{figure} We have performed a number of numerical solutions of the Fokker-Planck equation~(\ref{fund}), with the goal of validating the analytical approximations of Section~\ref{evol}. We use a finite difference code to examine the evolution of the electron velocity distribution $f(v,t)$ with time as governed by Equation~(\ref{fund}) with $D_{\rm turb}=D_0/v={\Gamma}/2\kappa v$. For the simulations, we adopted a typical value for $\kappa =5$, which agrees well with solar flare hard X-ray observations (cf. Equation~(\ref{kappa-gamma})). Firstly we check that we do indeed obtain a kappa distribution from the balance of Coulomb collisions and stochastic acceleration within Equation~(\ref{fund}). Figure~\ref{dist-evol-fig} shows the evolution of an originally Maxwellian thermal population of electrons (blue, solid line) toward a final state which agrees with the stationary solution kappa distribution (purple, dot-dot-dot-dashed line) as given by Equation~(\ref{kappa-dist}). We see that the distribution at $t = 100 \, \tau_{\rm c}$ closely approximates the kappa distribution form below $\simeq 5 \, v_{\rm te}$, corresponding to a range of about four orders of magnitude in $f(v,t)$. Such a distribution at $t=100 \, \tau_{\rm c}$ is thus close to a kappa distribution form for a significant proportion of the particles. \begin{figure}[htpb] \centering \includegraphics[width=10cm]{f4.eps} \caption{Evolution of the normalized distribution $f/f_{\kappa}$ with time. The solid blue line shows the normalized injected Maxwellian and then, from left to right: $f(v,t)/f_{\kappa}$ at $t/\tau_c = 10$ (orange dotted line), $t/\tau_{\rm c} =100$ (green dashed line), and $t/\tau_{\rm c}=300$ (red dot-dashed line). }\label{norm-dist-evol-fig} \end{figure} The evolution of the normalized distribution (Figure~\ref{norm-dist-evol-fig}) shows a ``wavefront'' moving towards higher energies, as expected from the advection-diffusion nature of Equation~(\ref{u-evol}). (We do not plot the final state of the distribution at $t = 1000 \, \tau_{\rm c}$ as it is almost a constant across the domain.) Examining $f/f_\kappa$ gives a clearer view of how close the electron distribution approximates a kappa distribution at different points of the simulation. Our results confirm that the electron distribution function at $t = 100 \, \tau_{\rm c}$ (green dashed line) is close to a kappa distribution for $v \lapprox 5 \, v_{\rm te}$ and that for $t = 300 \, \tau_{\rm c}$ it is almost indistinguishable from a kappa distribution up to around $7 \, v_{\rm te}$. In Section~\ref{evol} we found that the location of the front in velocity space evident in Figure~\ref{norm-dist-evol-fig} should depend on time according to $v_f(t) \sim t^{1/3}$ (Equation~(\ref{vf})). To assess the accuracy of this analytical result we arbitrarily choose a value $u(v,t) = f(v,t)/f_\kappa = 0.5$ to define the front location $v_f(t)$. A plot of $v_f$ versus time (in units of the collision time $\tau_{\rm c}$) is shown in Figure~\ref{vf-comparison-fig}. Before $t \simeq 20 \, \tau_{\rm c}$ there is a significant disagreement because the analytic expression~(\ref{vf}) holds only for $t \gg \tau_{\rm c}$, i.e., when a sufficient number of particles have been accelerated to non-thermal energies. At longer times $t \gapprox 700 \, \tau_{\rm c}$ a discrepancy also develops, which is due to the simulation results reaching the upper limit of velocity allowed in the system. However, for times between these two extremes, we see excellent agreement between the numerical and analytic solutions in terms\footnote{The constant offset between the curves in Figure~\ref{vf-comparison-fig} is not significant; it merely reflects the subjective nature of the choice $f(v,t)/f_{\kappa} = 0.5$ for the location of the velocity front.} of the power-law slope $d \ln v_f/d \ln t =1/3$. These numerical results show that the velocity-space front scenario as well as Equation~(\ref{vf}) provide a generally good description of the way particles are accelerated toward the kappa distribution in this model. \begin{figure}[htpb] \centering \includegraphics[width=10cm]{f5.eps} \caption{Location $v_f$ of the front in velocity space (in units of the thermal speed $v_{\rm te}$) versus time (in units of the collisional deceleration time $\tau_{\rm c}$). The analytic approximation for front speed $v_f(t)$ (Equation~(\ref{vf})) is shown by the orange solid line. The blue line shows the location of the velocity where $f/f_{\kappa}=0.5$, from numerical simulations.} \label{vf-comparison-fig} \end{figure} \section{STOCHASTIC ACCELERATION BY A LARGE SCALE ELECTRIC FIELD WITH STRONG PITCH-ANGLE SCATTERING}\label{stationary} The primary energy release in solar flares involves the reconnection of magnetic fields to produce electric fields. Various authors have considered the role of magnetic reconnection in particle acceleration, including large-scale sub-Dreicer \citep[e.g.,][]{1994ApJ...435..469B} and supra-Dreicer \citep[e.g.,][]{1993SoPh..146..127L,1996ApJ...462..997L} electric fields. Here we extend the analysis of large-scale coherent electric fields to include the role of turbulent pitch-angle scattering. A main objective of this analysis is to point out that efficient pitch-angle scattering of the particles in a region of constant electric field strength can still create an effect akin to stochastic acceleration, possibly suppressing the runaway phenomenon \citep{1994ApJ...435..469B} and preventing the production of an unacceptably large unidirectional current in these acceleration models. As we are interested in the formation of kappa distributions by turbulent acceleration we also discuss the conditions leading to a turbulent diffusion coefficient of the desired form $D_{\rm turb}(v)\sim v^{-1}$. Under the action of an accelerating electric field $E_\parallel$ (statvolt~cm$^{-1}$) parallel to the ambient magnetic field ${\bf B}$, the one-dimensional kinetic equation for a gyrotropic ($\partial f/\partial \phi=0$) distribution function $f(z,\beta, v,t)$ is \begin{equation}\label{kinetic} \frac{\partial f}{\partial t} + v \, \cos \beta \, \frac{\partial f}{\partial z} + \frac{eE_\parallel}{m_{e}} \, \mathbf{b}.\nabla_{\mathbf{v}}f = \frac{v}{\lambda} \, \frac{1}{\sin \beta} \, \frac{\partial}{\partial \beta} \left ( \sin \beta \, \frac{\partial f}{\partial \beta} \right ) \,\,\, , \end{equation} where $z$ (cm) is the position of the gyrocenter along the magnetic field with direction ${\mathbf b} = {\mathbf B_0}/B_0$, $\beta$ is the pitch angle ($\cos \beta = \mathbf {v}.\mathbf{B}_{0}/vB_{0} = v_{\parallel}/v$) and $v=\sqrt{v_{\parallel}^{2}+v_{\perp}^{2}}$ is the particle speed. In the case under consideration, the acceleration region is characterized by an electric field of constant magnitude $E_{\parallel}$ aligned with the direction $\mathbf{b}$ of the magnetic field. Transforming to the variables $(z,\mu,v,t)$, with $\mu= \cos \beta$, this may be rewritten as \begin{equation}\label{fokker} \frac{\partial f}{\partial t}+\mu \, v \, \frac{\partial f}{\partial z} + \frac{e E_\parallel}{m_{e}} \, \mu\, \frac{\partial f}{\partial v} +\frac{e E_\parallel}{m_{e}} \, \frac{(1-\mu^{2})}{v} \, \frac{\partial f}{\partial \mu} = \frac{v}{\lambda} \, \frac{\partial}{\partial \mu} \left [ (1-\mu^{2}) \, \frac{\partial f}{\partial \mu} \right ] \,\,\, . \end{equation} The last term in this equation describes pitch-angle diffusion, which tends to isotropize the distribution function on a (velocity-dependent) time scale given by \begin{equation}\label{taupa-def} \tau_{\rm pa}(v)=\frac{\lambda(v)}{v} \,\,\, , \end{equation} where the mean free path $\lambda(v)$ is generally a function of $v$. For instance, collisional pitch-angle scattering produces isotropization of the distribution function on time scale $\tau_{\rm pa}\sim v^{3}$ ($\sim v_{\rm te}^{3}$) corresponding to $\lambda\sim v^{4}$ ($\sim T^{2}$ for thermal particles) and the absence of an external electric field\footnote{for $\lambda \sim v^{4}$, Equation~(\ref{fokker}) is identical to the model studied by \citet{1964PhFl....7..407K} in the context of the formation of runaway electrons in plasmas, while the case of a velocity-independent mean free-path $\lambda \sim v^{0}$ corresponds to the standard Drude model of electric resistivity studied by Lorentz (1905).}. Acceleration of the particles is described by the third term in Equation~($\ref{fokker}$), i.e., \begin{equation}\label{dot-v} \dot{v}=\frac{e E_\parallel}{m_{e}} \, \mu \,\,\, , \end{equation} which shows that fluctuations in $\mu$ are also responsible for fluctuations in $v$. Given that particles are accelerated by the external electric field, it is quite clear that the isotropization effect of pitch angle scattering becomes dominant whenever $\tau_{\rm pa}(v) = \lambda(v)/v$ is a decreasing function of $v$. In fact, assuming $\tau_{\rm pa}(v)\sim v^{-\alpha}$, it can be shown that when $\alpha>0$ the particle distribution function remains close to isotropic despite the presence of the constant electric force \citep{1981JSP....24...45P,PhysRevE.56.3822,PhysRevLett.99.030601}. No runaway phenomenon occurs in this case. The combined effect of the electric field and pitch-angle scattering shows up as isotropic diffusive acceleration of electrons and an unlimited growth of their kinetic energy in the absence of collisional energy losses. The corresponding velocity-space diffusion coefficient can then be computed from the \citet{taylor} formula \begin{equation}\label{taylor-dturb} D_{\rm turb}(v) = \frac{e^{2}E_{\parallel}^{2}}{m_{e}^{2}} \int_{0}^{\infty} \langle \, \mu(0) \, \mu(t) \, \rangle \, dt =\frac{e^{2} \, E_\parallel^{2} \, \lambda(v)}{3 \, m_{e}^{2} \, v} \,\,\, . \end{equation} In the case where the mean free path $\lambda(v)$ is independent of $v$, this may be written \begin{equation}\label{taylor-dturb-one-over-v} D_{\rm turb}(v) = {D_{0} \over v} \,\,\, , \end{equation} with \begin{equation}\label{D0-result} D_{0} = \frac{e^{2} \, E_{\parallel}^{2} \, \lambda}{3 \, m_{e}^{2}} \,\,\, . \end{equation} As discussed in Section~\ref{kap}, in such a case the acceleration time $\tau_{\rm acc}(v) \sim v^2/D_{\rm turb}(v)$ (Equation~(\ref{tacc-def})) and the collisional deceleration time $\tau_{\rm c}(v)$ (Equation~(\ref{tdiff-def})) have the {\it same velocity dependence} ($\sim v^3$). Indeed, since $D_{\rm turb}(v) \propto \tau_{\rm pa}(v) = \lambda/v \sim v^{-1}$, Equation~(\ref{tacc-def}) shows that $dv/dt \sim v/\tau_{\rm acc} \sim v^{-2}$, so that $dE/dt \equiv m_e \, v \, dv/dt \sim v^{-1} \sim E^{-1/2}$; the kinetic energy thus grows like $E\propto t^{2/3}$ (see Equation~(\ref{vf})). The role of collisional energy losses is to allow the distribution of electrons to steadily converge toward the stationary kappa distribution~(\ref{kappa-dist}) as a result of the balance between turbulent acceleration and friction. In the above reasoning we have completely ignored the finite size of the acceleration region which imposes a maximum energy gain bounded by the finite electric potential drop across the acceleration region. Unfortunately, our treatment (Section~\ref{escape-analysis}) of escape from a finite-length acceleration region does not apply to the case of a stationary electric field -- whereas the maximum energy gained by particles from time-dependent electric fields in a finite length acceleration region depends on the confinement (escape) time $\tau_{\rm esc}$, the amount of energy gained by particles under the influence of a time-independent electric field is independent of the amount of time these particles stay confined in the acceleration region. Further, given that only particles moving parallel to the applied electromotive force $eE_\parallel$ gain energy, while those flowing antiparallel to the applied force lose it, a spatial asymmetry remains in such an acceleration model, despite the effects of isotropization. This is an undesirable feature of the model, requiring that some form of fragmentation of the electric field, such as oppositely directed electric fields on different magnetic field lines \citep[e.g.,][]{1985ApJ...293..584H,1995ApJ...446..371E,1997ApJ...489..367A,2004ApJ...608..540V,2008ApJ...687L.111B,2012SoPh..277..299G,2012SSRv..173..223C,2013SoPh..284..489G}, must be invoked. The value of $\kappa$ in the distribution~(\ref{kappa-dist}) is the ratio (Equation~(\ref{kappadef})) of the collision parameter $\Gamma$ (Equation~(\ref{gamma-def})) to the diffusion parameter $D_0$ (Equation~(\ref{dturb-result})) and hence, in a model involving stochastic acceleration by direct electric field, relates the ambient density $n$ to the (square of the) strength of the accelerating electric field $E_\parallel$ (Equation~(\ref{D0-result})). Thus the shape of the accelerated electron distribution constrains the values of one or both of these physical parameters. We now briefly explore the nature of this constraint as imposed by the observed shape of solar flare hard X-ray spectra. Using Equations~(\ref{kappadef}) and (\ref{D0-result}), we find that the value of the power-law index $\kappa$ in the distribution~(\ref{kappa-dist}) is given by \begin{equation}\label{kappa-dreicer} \kappa = \frac{\Gamma}{2D_{0}}=\frac{3}{2} \left ( \frac{\lambda _{c}}{\lambda} \right ) \left ( \frac{E_{D}}{E_{\parallel}} \right )^{2} \,\,\, , \end{equation} where we have introduced the usual collisional mean free-path \begin{equation}\label{lambda-coll} \lambda_{c} = \frac{(k_B T)^{2}}{4\pi ne^{4}\ln \Lambda} \end{equation} and the Dreicer field \begin{equation}\label{dreicer-def} E_D \equiv {k_B T \over e \lambda_c} = {4 \pi n \, e^3 \ln \Lambda \over k_B T} \,\,\, , \end{equation} i.e., the field strength required to accelerate an electron to the thermal energy over a distance equal to the collisional mean free path. As discussed in Section~\ref{kap}, observations of solar flare hard X-ray spectral shapes reveal that a typical value for $\kappa$ is $\kappa \simeq 5$ (Equation~(\ref{kappa-gamma})), which therefore provides the following constraint on the value of the accelerating electric field: \begin{equation}\label{e-constraint} E_{\parallel} \simeq \left ( \frac{3}{10}\frac{\lambda_{c}}{\lambda} \right )^{1/2} \, E_{D} \,\,\, . \end{equation} Moreover, normalizability of the kappa distribution~(\ref{kappa-dist}) requires that $\kappa > 3/2$, or $ E_{\parallel}< \left ( {\lambda_{\rm c} / \lambda } \right ) ^{1/2}E_{D}$. For typical conditions in the flaring loop-top source coronal plasma, $T \simeq 2 \times 10^7$~K and $n \simeq 10^{11}$~cm$^{-3}$, leading to a collisional mean free path $\lambda_c \simeq 5 \times 10^6$~cm and a Dreicer field $E_D \simeq 3 \times 10^{-4}$~V~cm$^{-1}$. Recently \citet{2014ApJ...780..176K} have argued, on the basis of the observed variation of hard X-ray source size with energy, that the turbulent mean free path $\lambda$ is in the range $10^8 - 10^9$~cm. Thus $\lambda_c/\lambda \simeq 0.005 - 0.05$, leading (Equation~(\ref{e-constraint})) to $E_{\parallel} \simeq (0.05 - 0.1)E_D \simeq (2 - 3) \times 10^{-5}$~V~cm$^{-1}$. Such a value of $E_{\parallel}$ is broadly consistent with the acceleration of electrons to deka-keV energies over observed loop lengths $L \simeq 10^9$~cm. \section{SUMMARY AND CONCLUSIONS}\label{conclusion} Driven by {\em RHESSI} observations of confined loop-top hard X-ray sources in solar flares, we have considered a model with cospatial stochastic acceleration, collisional deceleration and thermalization, and hard X-ray bremsstrahlung emission. For a turbulent diffusion coefficient associated with the acceleration mechanism of the form $D_{\rm turb} \sim 1/v$, and in the absence of particle escape, the electron distribution asymptotically approaches a kappa distribution~(\ref{kappa-dist}) with time. The approach toward this asymptotic steady-state kappa distribution proceeds as a ``wavefront'' in velocity space, with electrons of speed $v$ accelerated at successively greater times $t \sim v^3 \sim E^{3/2}$. This velocity-space front scenario, as well as the basic timescales involved, are supported by the results of numerical simulations. For sufficiently high velocities, the time taken to approach the kappa distribution becomes long enough that escape of electrons from the acceleration region can no longer be neglected. The effect of this was considered analytically in the limit of a small escape rate, when the acceleration region effectively behaves as a thick-target. With the high-spectral-resolution hard X-ray observations from {\em RHESSI}, the form of the hard X-ray-emitting (and hence, in this context, accelerated) electron distribution can be determined with impressive accuracy. Analysis of the spatially-integrated spectra from loop-top sources therefore provides a test of the predictions of the current model. Further, quantitative analysis of the energies, both low and high, at which the inferred electron distribution approaches and/or deviates from the asymptotic, escape-free, kappa distribution provides information on the value of the physical parameters of the model, such as the acceleration region length $L$ and the diffusion coefficient parameter $D_0$. \acknowledgments We thank the referee for drawing our attention to the different representations of the kappa distribution and its broader role in space plasma physics. This work is partially supported by a STFC grant. Financial support by the European Commission through the ``Radiosun'' (PEOPLE-2011-IRSES-295272) is gratefully acknowledged. AGE was supported by grant number NNX10AT78G from NASA's Heliospheric Physics Division. \bibliographystyle{apj}
\section{Introduction} Wide-field radio sky surveys yield information for large samples of Galactic and extra-galactic objects, allowing analysis of emission physics, source populations and cosmic evolution of radio sources such as active galactic nuclei, starforming galaxies, pulsars and supernova remnants. Surveys in new regions of parameter space, such as frequency, are particularly useful for identifying new and unusual objects and adding constraints to emission models. While radio surveys such as the Sydney University Molonglo Sky Survey \citep[SUMSS;][]{1999AJ....117.1578B,2003MNRAS.342.1117M} at 843\,MHz and the Molonglo Reference Catalogue \citep[MRC;][]{1981MNRAS.194..693L} at 408\,MHz cover the southern sky, there is no deep survey of the southern sky below these frequencies. The Culgoora Circular Array \citep{1995AuJPh..48..143S} measured approximately 1800 high-frequency-selected sources over a Dec range of $-48^\circ$ to $+35^\circ$ at 80~and 160\,MHz with limiting flux density of 4 and 2\,Jy, respectively. However, these were targeted observations rather than a survey. A recent low-resolution (26\hbox{$^\prime$\,}) survey by the Precision Array for Probing the Epoch of Reionization \citep[PAPER;][]{2011ApJ...734L..34J} detected $\approx500$ sources at 145\,MHz over 4,800\,deg$^2$. The new generation of aperture-array telescopes such as the Murchison Widefield Array \citep[MWA;][]{2013PASA...30....7T, 2009IEEEP..97.1497L} and the Low-Frequency Array \citep[LOFAR;][]{2013A&A...556A...2V} offer a chance to conduct all-sky low-frequency surveys at higher angular resolution, sensitivity, and speed than ever achieved before. Observations using the MWA 32-element prototype array such as \citet{2012ApJ...755...47W} and \citet{2013ApJ...771..105B} have demonstrated the scientific potential of low-frequency observations with the MWA design on a radio-quiet southern site. The MWA is the only low-frequency precursor\footnote{Defined as a facility exploring SKA technology, science and/or operations on an SKA candidate site: https://www.skatelescope.org/technology/precursors-pathfinders-design-studies/} for the Square Kilometre Array (SKA) and the first of the three SKA precursors to be operational for science. The MWA is located at the Murchison Radio-astronomy Observatory in outback Western Australia, the planned site of the future multi-million element SKA-low array. Thus the MWA explores the characteristics of the site at low frequencies in the pursuit of challenging science, exercising much of the physical infrastructure that will be applied to the SKA over a geographic distance of 800\,km (on-site infrastructure, long haul data transport, data archive infrastructure), and provides a valuable base for SKA verification systems. This paper presents the MWA Commissioning Survey (MWACS), work undertaken during the commissioning period of the full MWA, using a subset of its capabilities. The survey aimed to verify instrumental performance, verify our understanding of the MWA primary beam, motivate development of new data processing techniques and create an initial sky model of brighter radio sources at MWA frequencies. A comprehensive sky model, in particular, is a pre-requisite for calibrating the full MWA, due to its huge field-of-view \citep{2008ISTSP...2..707M}. The observed field also includes two of the three regions chosen for deep ($\approx1000$-hour) integrations for the MWA's Epoch of Reionisation key science program (see \citealt{2013PASA...30...31B} for a summary of the MWA's key science programs). MWACS is the first large-scale survey performed by the MWA in its role as an SKA precursor and the first such survey by an SKA precursor. These observations cover approximately {6,100} deg$^2$ roughly centred on the south Galactic pole. This is a region of sky with low brightness temperature, populated mostly by extragalactic sources, which we detect, catalogue and characterise in unprecedented detail at these frequencies. The MWACS is the first systematic exploration of the end-to-end MWA system and the characteristics of the SKA site at these frequencies, and lays a base of understanding for the much larger and more comprehensive GaLactic and Extragalactic All-sky MWA (GLEAM) survey, which is currently underway and will survey the entire sky south of $\delta=+20^\circ$ between 72 and 230\,MHz (Wayth et al., in prep). In turn, this will lay a base for the massive continuum surveys that will take place at low frequencies with the SKA, from the same site. This paper is laid out as follows. Section~\ref{sec:obs_datared} describes the observations, and the calibration and imaging strategy used in data reduction. We demonstrate how a hybrid strategy using conventional radio astronomy tools can generate high quality mosaics by taking advantage of the MWA's very good instantaneous $u,v$ coverage and near-coplanarity. Section~\ref{sec:sourcefindflux} describes the source-finding and flux density calibration procedures, including correcting for the MWA primary beam and how we have dealt with resolved structures. Section~\ref{sec:sourcecat} describes the properties of the source catalogue and discusses issues affecting reliability and sensitivity. Section~\ref{sec:conclusions} contains a discussion and conclusion. \section{Observations and data reduction} \label{sec:obs_datared} As detailed by \citet{2013PASA...30....7T}, the MWA consists of 128 32-dipole antenna ``tiles'' distributed over an area approximately 3~km in diameter. Each tile observes two instrumental polarisations, ``X'' (16~dipoles oriented East-West) and ``Y'' (16~dipoles oriented North-South). The zenith field-of-view of a beamformed tile is $\approx30^\circ$ at 150\,MHz. The signals from the tiles are collected by 16~in-field receiver units, each of which services eight tiles. For engineering reasons, during the commissioning period only four receivers were active at any one time, hence the tiles and receivers were commissioned as an overlapping series of six 32-tile sub-arrays labelled \emph{alpha} through \emph{zeta}. The sub-arrays were chosen to have good snapshot $u,v$-coverage within the various technical constraints in place during commissioning. The data presented in this paper were recorded by sub-arrays \emph{beta} and \emph{gamma}. The antenna layout and snapshot $u,v$-coverage of the combination of these two arrays are shown in Figure~\ref{fig:gamma}. The baselines are 8--1,530\,m in length, and their combined effective angular resolution at 180\,MHz is of order 3\hbox{$^\prime$\,} (5\hbox{$^\prime$\,} at 150\,MHz and 6\hbox{$^\prime$\,} at 120\,MHz.). For simplicity, only the ``XX'' and ``YY'' polarisations are used in the following analysis; the cross-polarisation terms are discarded. The effect of ignoring these terms is constant throughout the night, as the instrument gains are very stable; any small loss in flux is later fixed during the flux calibration stage (Section~\ref{sect:abscorrect}). \begin{figure*} \centering \begin{subfigure}[b]{0.3\textwidth} \label{fig:gamma_ants} \includegraphics[angle=-90,width=\textwidth]{Gamma_Beta_layout.png} \end{subfigure} \begin{subfigure}[b]{0.3\textwidth} \label{fig:gamma_ants_zoom} \includegraphics[angle=-90,width=\textwidth]{Gamma_Beta_layout_zoom.png} \end{subfigure} \begin{subfigure}[b]{0.3\textwidth} \label{fig:gamma_uv} \includegraphics[angle=-90,width=0.9\textwidth]{Gamma_Beta_uv_coverage.png} \end{subfigure} \caption{The left and middle panels show the antenna layout of \emph{beta} (empty squares) and \emph{gamma} (filled squares). The light gray shaded box on the left panel is enlarged in the middle panel, to more clearly display the central tiles. The monochromatic zenith-pointed snapshot combined $u,v$-coverage of the two MWA sub-arrays at 150\,MHz is shown on the right.} \label{fig:gamma} \end{figure*} A number of observation programs were undertaken during commissioning, one of which was night-time ``drift scans'' where the MWA tiles were pointed to a single Declination (Dec) along the meridian and data were collected as the sky drifted through the tile beams. This form of observation has previously been shown to be an effective way to observe large fractions of the sky with good sensitivity for the MWA \citep{2013ApJ...771..105B} and many other instruments, past and present, that use fixed dipole arrays also use drift scans. Using drift scans, uncertainty about the system calibration is minimised because the settings of all analogue components of the system are unchanged over the entire observation. The MWA has excellent stability of both the amplitude and phase of antenna gains using this mode, especially at night, when the ambient temperature changes slowly. We found that the standard deviation of primary-beam-corrected peak flux densities of typical bright unresolved ($>50$\,Jy) sources was only 1.5\% as they drifted through the zenith beam. Typical commissioning calibration scans show phase stability of better than than 1\arcdeg over the band. Overall, we found the quality of drift scan data to be limited by the stability of the ionosphere, which generates slow astrometric changes (see Section~\ref{sec:astrometry}); in calm conditions a single calibration solution can be applied to all the data for an entire night, in the same fashion as \citeauthor{2013ApJ...771..105B}. \begin{table*} \centering \caption{Summary of observations} \label{tab:datasummary} \begin{tabular}{lcccc} \hline Array & Central Dec & Date & Calibrator & Observing time / hours \\ \hline \emph{beta} & $-26.7^\circ$ & 2012 Oct 18 & 3C444 & 9.8 \\ \emph{beta} & $-47.5^\circ$ & 2012 Oct 19 & Pictor~A & 11.1 \\ \emph{gamma} & $-26.7^\circ$ & 2012 Oct 30 & 3C444 & 11.0 \\ \emph{gamma} & $-47.5^\circ$ & 2012 Oct 31 & Pictor~A & 11.0 \\ \hline \end{tabular} \medskip\\ $^a$All observations used three frequency settings centred on 119.04, 149.76 and 180.48\,MHz with 30.72\,MHz bandwidth.\\ \end{table*} The data presented in this paper are from drift scans taken at two Dec settings: $\delta=-26\fdg7$ (the zenith; hereafter referred to as Dec\,$-27$) and $\delta=-47\fdg5$ (hereafter referred to as Dec\,$-47$). The observations are broken into a series of scans that cycle between three frequency settings with centre frequencies 119.04, 149.76 and 180.48\,MHz (henceforth given as 120, 150 and 180\,MHz); the scans are two minutes long and are separated by eight-second gaps. After each scan, the frequency is changed and a new scan begins. The bandwidth of the MWA is 30.72\,MHz, hence the total frequency range of our observations is 104 to 196\,MHz. Observations at a particular frequency are thus separated into many two-minute scans that begin every six~minutes. In October, the night-time drift field-of-view encompasses Right Ascension (RA) in the range $\approx20$h--$09$h; we restricted our analysis to the range where both \emph{beta} and \emph{gamma} data were available at identical local sidereal times: $21\mathrm{h}<\mathrm{RA}<8\mathrm{h}$. The observations are summarised in Table~\ref{tab:datasummary}, and in total comprised 3\,TB of unaveraged visibilities (reduced to 540\,GB after flagging and averaging; see Section~\ref{subsect:flagandcal}). Figure~\ref{fig:drift-movie} shows an animation of snapshots of one frequency produced by a typical drift scan. Unlike \citet{2013ApJ...771..105B}, whose focus was the large-scale polarisation features of the low-frequency sky, these commissioning data are more suitable for measuring the flux densities of individual faint compact sources. \citeauthor{2013ApJ...771..105B} used the Real Time System \citep{2008ISTSP...2..707M,2010PASP..122.1353O} and a forward-modelling scheme to peel 250~sources from 2,400~square degrees of sky; given the instantaneous sensitivity of the commissioning sub-arrays, following such a procedure for our larger, more resolved sky, was infeasible. The following data reduction therefore uses standard astronomy tools such as Common Astronomy Software Applications (\textsc{casa}\footnote{http://casa.nrao.edu/}) and \textsc{miriad} \citep{1995ASPC...77..433S}. \begin{figure} \centering \includegraphics[width=0.5\textwidth]{image040.png} \caption{An animation, at four frames per second, showing the central $30^\circ\times30^\circ$ of the Dec\,$-47$ 180\,MHz drift data created by combining the 180\,MHz visibilities of the nights of 2012~Oct~19 and 2012~Oct~31, and following the imaging procedure outlined in Section~\ref{subsect:imaging}. In the usual convention, RA increases from right to left and Dec from bottom to top. The central Dec remains constant throughout at Dec$=-45^{\circ}35\arcmin35''$; the first frame is centred on RA$=21^{\mathrm{h}}26^{\mathrm{m}}57^{\mathrm{s}}$ and the last is centred on RA$=07^{\mathrm{h}}34^{\mathrm{m}}14^{\mathrm{s}}$. The colourscale is linear and runs from $-0.25$--1\,Jy\,beam$^{-1}$, but no correction has yet been made for the MWA primary beam. NB: Due to size limits, the animation is only visible in the published version of this article.} \label{fig:drift-movie} \end{figure} \subsection{Flagging and calibration}\label{subsect:flagandcal} The MWA band is comprised of 3072~10\,kHz `fine' channels, derived from 24~1.28\,MHz `coarse' channels. The eight edge fine channels of each coarse channel suffer aliasing and are flagged in every observation. The central fine channel of each coarse channel contains the (non-zero) DC component of the polyphase filterbank, so is also flagged. Known misbehaving antennas were flagged, and the Orbcomm transmission frequency range 136--138.5\,MHz was completely excised. Radio-frequency interference was excised using \textsc{aoflagger} \citep{2012A+A...539A..95O}; due to the radio-quiet nature of the observatory, less than three~per~cent of the data were affected and removed, mostly at the 136--138\,MHz range of the Orbcomm satellite downlink. From a native frequency resolution of 10\,kHz, the data were averaged by a factor of four in frequency, appropriately downweighting any frequency bin containing a reduced number of fine channels. From the original time-sampling of 0.5\,s, \emph{gamma} data were averaged by a further factor of two in time, while \emph{beta} data were time-averaged by a factor of eight, the difference arising from the larger number of long baselines in \emph{gamma}. A single calibration solution was determined for each frequency, for each night, based on the calibrator listed in Table~\ref{tab:datasummary}. \subsubsection{Dec\,$-27$ phase calibration: 3C444} The Dec\,$-27$ drift scans were phase-calibrated using 3C444: RA$=22^{\mathrm{h}}14^{\mathrm{m}}25.752^{\mathrm{s}}$; Dec$=-17^{\circ}01\arcmin36\farcs29$, which has a flux density $80$\,Jy and spectral index $-0.9$ at 160\,MHz \citep{1995AuJPh..48..143S}. 3C444 dominates the visibilities when observed near the meridian, and is unresolved at the resolution of the commissioning array. A two-minute observation produces solutions of S/N$\simeq5$, while integrating over the three two-minute snapshots in which 3C444 was closest to the meridian increases the S/N to $\simeq7$. As the sky has drifted by a total of 18\,minutes by the time data have been taken at all three frequencies, the instrumental response is different for each snapshot, so the expected flux of 3C444 was scaled by a primary beam model for each snapshot (see Section~\ref{sect:abscorrect} for more details on the primary beam model). A least-squares fit over the full duration of each observation was performed on those visibilities with $u,v$-distance\,$>0.1$\,k$\lambda$ using \texttt{bandpass}, a \textsc{casa} routine, to produce a single complex gain solution for each tile, for each polarisation, for each frequency interval, which was then applied to the night's observations. \subsubsection{Dec\,$-47$ phase calibration: Pictor~A} The Dec\,$-47$ drift scans were calibrated using Pictor~A: RA$=05^{\mathrm{h}}19^{\mathrm{m}}49.7^{\mathrm{s}}$; Dec$=-45^{\circ}46\arcmin43\farcs70$. Pictor~A was used only as a phase calibrator; its flux density of $S\simeq400$\,Jy \citep{1995AuJPh..48..143S} over the MWA observing band dominates the visibilities such that a single two-minute observation is sufficient to obtain per-channel solutions with S/N$\gg20$. Since Pictor~A is resolved by our instrument, the Very Large Array (VLA) images at 1400 and 333\,MHz \citep{1997A+A...328...12P} were used to extrapolate a spatial and spectral model for each MWA observing frequency\footnote{While the VLA 74\,MHz image is closer to the MWA observing band, it possesses insufficient North-South resolution to accurately model the source structure, and therefore provide a model to properly phase-calibrate the MWA visibilities.}. The same procedure was followed as with 3C444: the model was Fourier-transformed and used to generate a calibration solution via \texttt{bandpass}, which was then applied to the night's observations. After calibration, the \emph{beta} and \emph{gamma} data were concatenated. We note that the ionosphere was different between the two nights, but only the \emph{gamma} data posseses baselines of sufficient length to experience phase offsets significant compared to the width of the synthesised beam. Therefore it is possible to combine the \emph{beta} and \emph{gamma} visibility data for the same area of sky, but not to combine all of the \emph{gamma} data together, primary beam effects aside. \subsection{Subtraction of bright sources in primary-beam sidelobes} The first sidelobes of the MWA primary beam are estimated to have a response of approximately 3\% of that of the main beam, and the second around 0.1\% (see Section~\ref{sect:abscorrect} for more details on the primary beam model). When an extremely bright source such as Cygnus~A passes through such a sidelobe, it contributes signal to the visibilities. If not subtracted directly from the visibilities (``peeled'') or deconvolved, this contribution is visible in the image as striping from the sidelobes of the synthesised beam at the position of the source. In the absence of computing constraints, one could attempt to image the entire sky, such that all sources are deconvolved. However, we found that bright sources sufficiently far ($>45^{\circ}$) from the phase centre were poorly deconvolved even in a snapshot image due to the array not being perfectly coplanar. At the time of analysis, techniques such as $w$-projection \citep{2008ISTSP...2..647C} were too computationally expensive for the $\sim$10,000~pixel-wide grid required to cover the sky. The sidelobes also introduce a large spectral variation to the image pixels, as they vary much more quickly with frequency than the main beam. To reduce the effects of bright sources in the sidelobes of the primary beam, we phase-rotated the calibrated, concatenated snapshot datasets to the source positions and performed a shallow (threshold\,$=10$\,Jy) \textsc{clean} of a $1^{\circ}\times1^{\circ}$ around each source, with an additional Taylor term, allowing each pixel to have its own spectral index \citep{1994A+AS..108..585S}. The model pixels were then subtracted from the visibilities, which were then phase-rotated back to their original positions, ready for imaging. The sources removed in this way were Pictor~A, the Crab Nebula, Hydra~A, Cygnus~A, PKS$2153-69$ and PKS$2356-61$. Only Pictor~A was present in the primary beam main lobe, as well as the sidelobe, and we note that this strategy means we are then unable to recover its flux density in the final mosaic. \subsection{Imaging strategy}\label{subsect:imaging} At 150\,MHz, the full-width half maximum (FWHM) of the MWA primary beam at zenith is approximately $30^\circ$, corresponding to more than two hours of RA. Sources therefore appear in many of the two~minute scans as the sky drifts through the beam. Within any two-minute scan (hereafter a ``snapshot''), the hybrid \emph{beta}/\emph{gamma} array is sufficiently coplanar that snapshot images do not suffer from wide-field effects within the main primary beam lobe ($<15^\circ$ from the phase centre) and the image coordinates can be described by a slant-orthographic (generalised SIN) projection \citep{1999ASPC..180..383P,2002A+A...395.1077C}. The data were thus divided into snapshots with the phase centre defined as the local sidereal time in the middle of the snapshot. The goal of data processing is to form an image of the entire RA range of the observation using all of the available data. Given the drift scan observation strategy, standard synthesis imaging techniques are not suitable. In a drift scan, instead of pointing the beam at adjacent locations on the sky, the beam stays constant and the sky moves, relatively. The data are therefore best processed using a mosaicing methodology. Typically the goal of a mosaic observation is to image objects larger than the primary beam of the telescope \citep[e.g.][]{1999ASPC..180..401H}. Much of the mosaicing literature and software is therefore focused on recovering the large-scale structure in a synthesis image using an interferometer with very short baselines. In the case of these data, the commissioning sub-array has only 12 baselines shorter than 20\,m and thus sensitive to scales $>5\arcdeg$, unlike the full MWA which fully samples up to scales of $12\arcdeg$ (at 180\,MHz). As we are also observing a part of the sky mostly devoid of objects of large angular scale, our focus is on combining the data to maximise sensitivity to compact structures. To form the mosaic images, we perform the following steps, for each frequency band, for each polarisation: \begin{itemize} \item generate a output accumulation image for the data and primary beam in an equal area coordinate system with 0\farcm5 pixel resolution and over a sufficient region of the sky for all data; \item for each snapshot: \begin{itemize} \item invert using multi-frequency synthesis over a field-of-view of $40^{\circ}\times40^\circ$, with a pixel resolution of 1\hbox{$^\prime$\,}, weighting each $u,v$-cell equally (``uniform weighting''); \item \textsc{clean} to a threshold of three times the typical snapshot RMS, which itself is typically $\simeq240$--80\,mJy\,beam$^{-1}$ from 120--180\,MHz; the synthesised beam sidelobes are only $\approx10\%$ of the peak response, and a gain factor of $0.1$ is used, thus avoiding \textsc{clean}ing the noise, and \textsc{clean} bias; \item restore using an elliptical Gaussian approximation of the snapshot synthesised beam; \item regrid and add the deconvolved image multiplied by the appropriate primary beam model into the accumulation data image; \item likewise, regrid and add the square of the beam image into the accumulation beam image; \end{itemize} \item after all additions, divide the data image by the beam-squared image to form the final mosaic. \end{itemize} In essence, this follows equation (1) from \citet{1996A+AS..120..375S} which maximises the signal-to-noise ratio (S/N) in the output mosaic given the changing S/N over the field in each snapshot due to the primary beam. We use the \textsc{miriad} for the imaging, deconvolution and image arithmetic, and use our own code to generate MWA primary beam models (see Section~\ref{sect:abscorrect} for more details on the primary beam model). \textsc{miriad}'s understanding of advanced World Coordinate Systems (WCS) and ability to regrid between arbitrary coordinate frames is especially useful. \begin{figure*} \centering \includegraphics[width=0.8\textwidth]{example_mwacs_image_aplpy.png} \caption{A randomly-chosen section of the 180-MHz Dec\,$-47$ pseudo-Stokes~I mosaic, representing about 4\% of the total MWACS survey area (in one of three frequency bands). The greyscale is linear and runs from $-0.2$ to $+0.8$\,Jy, and the estimated PSF at the centre of the image is shown as a boxed filled ellipse at the bottom-left of the image, of dimensions $4\farcm31\times2\farcm94$, position angle $-50\fdg4$. Some calibration errors are still in evidence, particularly around the bright source in the North-West. MWACS detections are shown as white ellipses of the same shapes as their fitted Gaussian parameters. The RMS of this image is 28\,mJy.} \label{fig:example-mosaic} \end{figure*} The process of adding snapshots improves the quality of the final image by reducing the thermal noise and improving the synthesised beam, revealing many fainter sources. These fainter sources have not been deconvolved in the snapshots, so their sidelobes remain in the mosaic with $\sim$1~hour of effective earth rotation synthesis (the time it takes for sources to move through the utilised area of the primary beam). We estimate the ($5\sigma$) classical confusion limit \citep{1974ApJ...188..279C} to be approximately 5\,mJy\,beam$^{-1}$ given our effective synthesised beam size; this is consistent with the confusion measurement by \citet{2010A+A...522A..67B,2009A+A...500..965B} at 150\,MHz at a similar angular resolution and frequency. While we therefore begin to approach the classical confusion limit, the dominant components of noise in our mosaics are sidelobe confusion and calibration errors. A randomly-chosen section of the 180-MHz Dec\,$-47$ mosaic is shown in Figure~\ref{fig:example-mosaic}, showing the well-behaved point spread function (PSF) and averaging down of calibration errors that results from the mosaicking process. The zenith-pointed MWA primary beam is virtually identical for the XX and YY polarisations. We therefore averaged the XX and YY Dec\,$-27$ mosaics to form a (pseudo) Stokes-I image. For the Dec\,$-47$ mosaic, we found that the primary beam response of the XX and YY polarisations was sufficiently different that Dec-dependent corrections were required. Hence, the XX and YY mosaics for Dec\,$-47$ were not combined at this stage. Details of this process and absolute calibration are discussed in Section~\ref{sect:abscorrect}. \begin{table} \centering \caption{Gaussian parameters for corrected PSFs for the mosaics, at their original phase centres. The Dec\,$-47$ entries apply to both the XX and YY mosaics.} \label{tab:psf} \begin{tabular}{ccccc} \hline Dec & $\nu$/MHz & $a$/' & $b$/' & $\theta$/$^{\circ}$\tabularnewline \hline -27 & 119 & 6.17 & 4.30 & -63.8 \tabularnewline -27 & 150 & 5.25 & 3.30 & -63.7 \tabularnewline -27 & 180 & 4.41 & 2.79 & -63.6 \tabularnewline -47 & 119 & 6.44 & 4.45 & -50.3 \tabularnewline -47 & 150 & 5.08 & 3.47 & -50.4 \tabularnewline -47 & 180 & 4.31 & 2.94 & -50.4 \tabularnewline \hline \end{tabular} \end{table} \section{Source finding and flux-density estimation} \label{sec:sourcefindflux} The imaging process results in nine mosaics, consisting of three frequencies for the Dec\,$-27$ scan, and three frequencies $\times$ two polarisations for the Dec\,$-47$ scan. The RMS noise level of these mosaics is typically a factor of five lower than that of the individual snapshots, allowing much deeper source-finding; see Section~\ref{sect:noise} for a discussion of the origins of noise in the mosaics, and their effect on the accuracy of our source flux densities. \subsection{Sensitivity} \begin{figure*}[t] \centering \begin{subfigure}{\textwidth} \centering \includegraphics[width=0.9\textwidth]{C102_141_rms.png} \label{fig:c102_rms} \end{subfigure} \begin{subfigure}{\textwidth} \centering \includegraphics[width=0.9\textwidth]{C103_141_rms.png} \label{fig:c103_rms} \end{subfigure} \caption{Log-scaled images of the root-mean-square (RMS) intensity measured across the mosaics for the highest-frequency Dec\,$-27$ (top) and Dec\,$-47$ (bottom) scan. The greyscale runs from 0.02 to 0.4~Jy\,beam$^{-1}$. The North--South axis is Dec in decimal degrees, and the East-West axis is RA in hours.} \label{fig:rmsmap} \end{figure*} Maps of the root-mean-square (RMS) value of the 180\,MHz mosaic pixels were generated using \textsc{Aegean} \citep{2012MNRAS.422.1812H} which calculates the RMS for blocks of pixels with dimensions of approximately 20$\times$20 beams\footnote{The RMS is calculated as the Inter-Quartile Range scaled by 1.34896. This is equivalent to the RMS for a Gaussian distribution but is more robust against high-flux pixels from bright sources \citep[c.f. e.g.][Section~5.2.1]{1998AJ....115.1693C}.}. These are reproduced for two representative mosaics in Figure~\ref{fig:rmsmap} (see also Figure~\ref{fig:rms_cdf}). Two effects can clearly be seen in these figures: an increase in RMS at the edges of the mosaics, and isolated patches of locally high RMS. The first is due to the effect of the primary beam and the second is due to the effect of sidelobe confusion and calibration errors around the brighter sources. The latter is particularly severe near extremely bright, resolved structures which can most easily be seen close to the Galactic plane (at the Easternmost edge of the Dec\,$-47$ scan). Fornax~A produces more calibration-error noise in the Dec\,$-27$ scan, and less in the better-calibrated Dec\,$-47$ scan; deconvolution errors from poorly reconstructing the largest scales of its structure are present in both. \subsection{Source identification}\label{subsect:source-ident} We wish to identify and characterise the morphology and flux density of all sources within the field which are bright enough to be discerned from the background noise. Having characterised the noise properties of the mosaics, we identify those pixels of the mosaics whose flux is greater than five times the local RMS, which itself is hereafter denoted $\sigma$. \textsc{Blobcat} \citep{2012MNRAS.425..979H} was used to identify `blobs': islands of pixels containing a $5\sigma$ peak, expanded to include all contiguous pixels brighter than $3.5\sigma$, while accounting for various biases. This was repeated for all three (Dec\,$-27$, Dec\,$-47$ XX and Dec\,$-47$ YY) of the highest frequency (180\,MHz) mosaics. These were used exclusively for source-finding since the highest resolution allowed the greatest separation of neighbouring sources into discrete blobs. The objects detected in the three maps were combined into a single list (i.e. from this point on no regard was taken as to which map the source was detected in). \subsection{Characterising the point spread function} For an interferometer coplanar to the plane perpendicular to the zenith, observing at zero hour angle, the synthesised beam becomes elongated N-S relative to the zenith, as $\csc\zeta$, where $\zeta$ is the zenith angle. Regridding from a slant orthographic projection to a zenithal equal area projection (using standard \textsc{miriad} WCS tools) conserves the synthesised beam volume, ignoring this effect, such that a correction factor of $\csc\zeta$ must be applied to rescale the N-S extent of the synthesised beam. We apply this immediately, before making further measurements. While the synthesised beam shape for an individual snapshot is known very accurately (short-timescale ionospheric effects aside), determination of the effective synthesised beam in the mosaics is a much more difficult problem. The mosaic contains approximately an hour (depending on frequency) of effective earth rotation synthesis which significantly circularises the PSF. However, refraction from the ionosphere acts differently in each snapshot and enlarges the PSF as snapshots are combined to form each mosaic. To measure this effect, we fitted 2D Gaussians to all detected sources, and plotted the distribution of ellipse values against various parameters (such as RA, Dec) and checked for correlations: none were seen. Plotting against S/N, as in Figure~\ref{fig:SNR-a_funnel_plot}, shows that the ellipse sizes of unresolved sources form a line, with the resolved sources having larger values. For each Gaussian parameter, a line was fit to the unresolved sources, and compared with the predicted value in order to calculate the correction needed to increase the size of the PSF to match the data. Henceforth, all source-finding uses the corrected PSFs, shown in Table~\ref{tab:psf}. Note that the PSFs are identical for the XX and YY polarisation Dec\,$-47$ mosaics, as all of these effects are polarisation-independent. A further projection factor of $\csc\zeta$ is applied when measuring sources in the mosaics. Figure~\ref{fig:beam_stretch} shows an exaggerated illustration of the effect of these changes. After these corrections, it is possible to recover the flux densities of extended sources, which would otherwise be over- or under-estimated depending on the difference between the local PSF and average PSF.\footnote{We note some similarity in this process to typical optical image processing, in which the PSF is determined by measurements of unresolved stars.} \begin{figure} \centering \includegraphics[angle=-90,width=0.5\textwidth]{S-N_vs_major_axis_180MHz.png} \caption{An example of the analysis used to find the correct PSF for the mosaicked images: in this case, we examine the major axis of sources detected in the zenith scan at 180\,MHz. The black crosses show the measured major axis of each source against its signal-to-noise ratio (S/N); the grey stars show the expected major axis if the zenith-angle-dependent projection effect were the only source of change in the synthesised beam; the light grey dashed line shows a S/N-weighted horizontal fit to the major axis measurements. The ratio of this fit to the predicted major axis gives the correction factor by which the major axis of the PSF must be increased to match the data: see Table~\ref{tab:psf} for a list of the corrected PSFs for each mosaic.} \label{fig:SNR-a_funnel_plot} \end{figure} \begin{figure} \centering \includegraphics[width=0.5\textwidth]{beam_compare.png} \caption{An exaggerated example of the corrections made to the PSF: the solid black ellipse shows a 150\,MHz synthesised beam at the zenith; the dashed line shows the synthesised beam for a pointing due South, at a zenith angle of $45^{\circ}$ (15$^\circ$ further than the maximum zenith angle of MWACS), resulting in a stretch in Dec; the dotted line shows an increase in the size of the PSF by a further 5\%; the real magnitude of the correction is closer to 2\%, and accounts for the effects of ionospheric smearing and image-based mosaicking.} \label{fig:beam_stretch} \end{figure} \subsection{Characterisation of discrete sources}\label{subsect:sourcechar} We now aim to determine flux density and morphology as accurately as possible given the three (Dec\,$-27$), six (Dec\,$-47$) or nine (overlapping region) measurements that we have for each source. On visual inspection it is clear that the vast majority of our objects can be characterised by a single elliptical Gaussian similar in shape to the synthesised beam. This reflects the fact that the vast majority of our sources are unresolved, as expected. In order that they be fit with the minimum number of free parameters possible, we begin by modelling each of our sources as a single elliptical Gaussian, only attempting to characterise more complex morphology when this assumption is determined to be invalid (see Section~\ref{sect:goodbadsplit}). In another refinement to avoid the effects of confusion due to nearby sources, the fits were performed only over those pixels which lie within the FWHM of the \textit{a priori} fit (with an extra margin of 10\%). Levensburg-Macquart least-squares fits were made to these pixels, with free parameters: amplitude; major axis, $a$; minor axis, $b$; position angle, $\theta$; RA offset, $\Delta_\mathrm{RA}$; and Dec offset, $\Delta_\mathrm{Dec}$. The initial parameters were the synthesised beam (with frequency and position dependence calculated as described in Section~\ref{subsect:imaging}) with an amplitude of the peak pixel in the region of the fit, centred on the flux weighted centroid of the object. For error calculation purposes the RMS in the region of the blob in each map was determined by taking the mean of the RMS map over an array of $24\times24$ pixels centred on the centroid of the blob. This provides a modest amount of smoothing of the RMS shown in Figure~\ref{fig:rmsmap} (where the RMS is determined over blocks a factor of $\sim5\times5$ larger). \subsubsection{Separating single-component and multi-component sources} \label{sect:goodbadsplit} Next we determine which of the blobs are indeed well-characterised by a single elliptical Gaussian and which require further characterisation. Our criterion is that any source whose centroid lies within the half-maximum of the \textit{a priori} elliptical Gaussian is ``single component''. This simple criterion is effective due to our selection of the \textit{a priori} centroid on the basis of the amplitude-weighted centroid of the blob. When there is no peak (i.e. positive curvature in both spatial dimensions) within our stringent pixel range, the centroid of the fit is forced outside the \textit{a priori} ellipse. If there is a peak, but it is not coincident with the \textit{a priori} centroid, this is indicative of signal within the blob which is not associated with the fitted Gaussian. Using this method, {12,863} ($\approx91$\,\%) of our {14,110} sources are classified as single-component. \subsection{Flux-density calibration} \label{sect:abscorrect} The southern sky at low frequencies is poorly surveyed and flux densities derived from extrapolation of various radio catalogues to the MWA frequency range disagree at the 10--20\% level. Work is under way in the community to develop a unified radio flux density scale from MHz--GHz frequencies\footnote{http://mwsky.ncra.tifr.res.in/mwsky/upload\_talk/11-Dec\_3B\_Rick\_1337\_y.pdf}, and GLEAM will expand this scale over the entire Southern sky. At the time of writing, the best course of action was to bootstrap from the fairly well-understood Northern sky to the South. Conveniently these declinations pass close to the zenith for the MWA, allowing the use of a simple model of the MWA primary beam. The MWA primary beam has previously been approximated by an analytic model incorporating a dipole over groundscreen and geometric array factor with good results \citep[e.g.][]{2012ApJ...755...47W,2013MNRAS.436.1286M,2014MNRAS.438..352B}, and we adopt the same approach in this paper. \citet{2013ApJ...771..105B} find that the zenith primary beam model is incorrect at the $\sim$2\% level at higher frequencies, and we expect the model to be less correct further from zenith as the simple dipole approximation becomes less valid. Due to the uncertainty in the primary beam model, especially away from the zenith, it is impossible to correctly perform the absolute flux density calibration in a single step. In addition, phase errors from imperfect phase calibration will reduce the recovered peak flux of radio sources, necessitating a small, position-independent correction. We note that understanding and improving the primary beam model is an ongoing activity within the MWA collaboration. Recent work \citep{2014Sutinjo} to incorporate mutual coupling effects into the model has improved our understanding of the response of the MWA beam. However the work is still ongoing and for the purposes of this paper our empirical method to bootstrap the flux density scale, described below, was sufficient. \begin{figure} \centering \includegraphics[width=0.9\columnwidth,clip=]{3c32_nvss.png} \caption{A postage stamp image of 3C32 extracted from the NVSS survey. Peak flux density is 2.11\,Jy\,beam$^{-1}$. Contours begin at 0.2\% peak and have a common ratio of 2.} \label{fig:3c32_nvss} \end{figure} \begin{figure} \centering \includegraphics[width=1.1\columnwidth]{0105-163.png} \caption{3C32 flux densities from the VLSS, Culgoora, MRC, and NVSS surveys shown with uncorrected MWA fluxes. Two fits to the non-MWA points are shown: the solid line is a power-law (i.e. a straight-line fit in log-log space); the dashed line fits a parabola in log-log space. MWA flux densities were corrected to lie on the power-law fit. } \label{fig:3c32_spectrum} \end{figure} In order to provide an absolute flux density scale, a list of sources was drawn up which appear in our catalogue as well as all of the Culgoora \citep[80, 160\,MHz][]{1995AuJPh..48..143S}, VLSS \citep[74\,MHz][]{2007AJ....134.1245C}, WISH \citep[325\,MHz][]{2002A&A...394...59D}, MRC \citep[408\,MHz][]{1981MNRAS.194..693L} and NVSS \citep[1.4\,GHz][]{1998AJ....115.1693C} catalogues. A good choice of flux calibrator might be our northern phase calibrator, 3C444; however, there are two reasons to avoid using this source. Firstly, it is towards the edge of our RA range, and is only sampled by a few snapshots. Secondly, it is resolved into two components by NVSS, so measurements by different instruments with different resolutions may give conflicting results. The second brightest candidate is 3C32. This source as it appears in NVSS (see Figure~\ref{fig:3c32_nvss}) would be unresolved in our data; its flux density is well-modelled by a simple power law (Figure~\ref{fig:3c32_spectrum}) and shows no evidence of variability or being resolved. In particular, the VLSS 74\,MHz and Culgoora 80\,MHz flux densities both lie on the power-law fit within the errors, despite their differing resolutions and epochs of observation. We performed an image-plane Gaussian fit (Figure~\ref{fig:3c32_pixmap}) to measure the flux density of the source and confirmed it as unresolved. We therefore scaled the integrated flux densities of each of the Dec\,$-27$ maps appropriately to fit the least-squares fit to the catalogue fluxes (see Table~\ref{tab:abscorrect}). Since the zenith analytic primary beam model had already been divided out in creating the Dec\,$-27$ mosaic, the relative flux densities of sources were already correct and only this single scaling factor should be required to set the absolute flux density scale across the whole mosaic, assuming that the zenith primary beam model is correct. \begin{table} \centering \caption{Scaling factors applied to the measured flux densities. The Dec\,$-27$ corrections make the fluxes consistent with the absolute flux scale determined for 3C32. The Dec\,$-47$ corrections make the flux densitites in the two fields consistent with each other.} \label{tab:abscorrect} \begin{tabular}{lcc} \hline $\nu$/MHz & Dec\,$-27$ Correction & Dec\,$-47$ Correction \\ \hline 119 & 0.904 & 0.600 \\ 150 & 0.963 & 0.769 \\ 180 & 0.939 & 0.818 \\ \hline \end{tabular} \end{table} The flux densities measured in the XX and YY mosaics of the Dec\,$-47$ field disagree at the $\sim$5\% level due to unmodelled primary beam effects. A second-order polynomial fit (weighted by the two flux densities added in quadrature) was made to the Dec-dependent discrepancy between XX and YY as shown in Figure~\ref{fig:xx_yy}. The XX mosaic was then scaled to match the YY mosaic. The Dec\,$-47$ mosaic was corrected to be consistent with the Dec\,$-27$ mosaic by using the $\approx600$ unresolved sources which lay in the overlapping region. The mean of the flux density ratio between the two maps of all of these sources was taken to be the additional correction factor to the Dec\,$-47$ mosaic to produce a corrected absolute flux density calibration scale consistent with that used for Dec\,$-27$ (Table~\ref{tab:abscorrect}). After this correction, the XX and YY mosaics of the Dec\,$-47$ field were combined, weighted by their respective primary beam responses (which differ by around 5\%), to form a pseudo-Stokes-I mosaic, on which source-finding can be performed. \begin{figure*} \centering \rotatebox{90}{ \begin{minipage}{\textheight} \begin{subfigure}[b]{0.3\textwidth} \label{fig:3c32_93} \includegraphics[width=1.0\textwidth]{3C32_93.png} \end{subfigure} \begin{subfigure}[b]{0.3\textwidth} \label{fig:3c32_117} \includegraphics[width=1.0\textwidth]{3C32_117.png} \end{subfigure} \begin{subfigure}[b]{0.3\textwidth} \label{fig:3c32_141} \includegraphics[width=1.0\textwidth]{3C32_141.png} \end{subfigure} \caption{\label{fig:3c32_pixmap} Least-squares fits to pixel values for the source 3C32 at 119, 150 and 180\,MHz. RA and Dec are in degrees; axes of top figure are arcminutes. Circles represent image pixels with circle diameter proportion to brightness. Unfilled pixels are those with at least 50\% of the brightness of the \textsc{Blobcat} peak. Grey are those with at least 25\% of the \textsc{Blobcat} peak. Red are those with negative values. Solid ellipses show half-power-beam-width and $\sqrt{2}\times$half-power-beam-width of the fitted elliptical Gaussian. Dashed ellipses show the synthesised beam. Bottom-left plots and residuals show fit of pixels with $a$, $b$ and $\theta$ constrained to the synthesised beam parameters. Bottom-right plots and residuals show fit of pixels with $a$, $b$ and $\theta$ as free parameters. Vertical lines on bottom plots correspond to the ellipses on the upper plot. The fit ellipses are barely discernible from the synthesised beam ellipses, indicating that the sources is unresolved. } \end{minipage} } \end{figure*} \begin{figure \centering \includegraphics[width=\columnwidth]{xx_yy.png} \caption{Ratio of XX to YY flux density for all sources detected in both maps plotted against Dec for all frequencies (low to high). The colour axis is the log$_{10}$ of the source flux density in Jy (as measured in the YY scan before amplitude calibration).} \label{fig:xx_yy} \end{figure} \subsection{Combining the mosaics} Many MWACS sources are detected both in the Dec\,$-27$ and Dec\,$-47$ mosaics. Combining the information from both mosaics is likely to give superior results for a number of reasons. This would recover some of the signal-to-noise in the overlapping region of the two mosaics (at Dec$\simeq-36^\circ$) which is otherwise attenuated by the primary beam, especially at the highest frequency. The differing primary beam between the two scans also means that the signal from sources in the distant sidelobes of the primary beam will be somewhat different, so combining the mosaics would reduce this noise. Unfortunately, the very different state of the ionosphere between the two \emph{gamma} drift scans makes a simple image-plane combination difficult; without extensive modelling of the ionosphere, there would be large position- and frequency-dependent smearing of the sources. This is evident from Figure~\ref{fig:astrometry}; the astrometry errors are time-dependent (seen as a change in RA) and differ night-to-night. However, an improved flux-density estimate can be determined for each source in the overlapping region by taking the mean of the two peak flux densities, weighted by the S/N of the fit in each of the maps (as defined in Section~\ref{sect:errors}). Though the size of the overlapping region varies dramatically with frequency, for simplicity and transparency the map flux densities were only combined in the region well-sampled by both the Dec\,$-27$ and Dec\,$-47$ scans at the highest frequency: a Dec range of $-38\fdg20 < \delta < -35\fdg25$. Furthermore, the positions and morphologies of the sources in the overlap region were taken as those measured in the Dec\,$-47$ mosaic, as this scan has a somewhat more compact synthesised beam and marginally better phase calibration (see Section~\ref{sect:noise}). \subsection{Fitting 180\,MHz flux densities} The majority of the MWACS source-flux densities are well-characterised by a simple power-law spectral index; as described in Section~\ref{subsect:source-ident}, we perform source detection at the highest resolution available, at 180\,MHz, so we use the fitted integrated source flux densities at each frequency to measure the spectral indices. Alongside this, we report the 180\,MHz flux density for each source. We performed weighted least-squares fits to the flux densities at the three frequencies, propagating the confusion and fitting errors described in Section~\ref{sect:errors}. \subsection{Estimation of errors} \label{sect:errors} \subsubsection{Noise-like errors}\label{sect:noise-like-errors} \citet{1997PASP..109..166C} addresses the problem of determining the error on the parameters of a least-squares fit of an elliptical Gaussian $G(x_k, y_k)$ to a set of pixel values $a_k$. Here, we adopt the same notation where $A$, $\theta_M$, $\theta_m$, $\phi$ $x_0$ and $y_0$ are the amplitude, FWHM major and minor axes, position angle and centroid offsets, respectively, of a Gaussian fit over $a_k$ pixel amplitudes, each having the same Gaussian error distribution of RMS $\mu$. In our case, the noise, whether it be thermal or confusion noise, will be correlated on the scale of the synthesised beam. For the case of unresolved or weakly resolved sources, where the correlation of the noise matches the apparent source size, the error on $A^2$, $\mu^2(A)$ is $\mu^2$, i.e. the error on the flux density of the source due to noise is simply the local RMS. The errors on $A$, $\theta_M$ and $\theta_m$ are therefore: \begin{equation} \label{eqn:err_fit} \frac{\mu^2(A)}{A^2} = \frac{\mu^2(\theta_M)}{\theta_M^2} = \frac{\mu^2(\theta_m)}{\theta_m^2} = \frac{\mu^2}{A^2} . \end{equation} By analogy to \citet{1997PASP..109..166C} equation~21, \begin{equation} \mu^2(\phi) = 2\left(\frac{\theta_M\theta_m}{\theta_M^2-\theta_m^2}\right) \frac{\mu^2(A)}{A^2} . \end{equation} Where $\mu(\phi) > 90^{\circ}$, $\mu(\phi)$ is set to null. Finally, the errors on RA ($\alpha$) and Dec ($\delta$) due to noise-like errors are given by \begin{eqnarray} \mu^2(\alpha) &=& \mu^2(\theta_M)\sin^2(\phi) + \mu^2(\theta_m)\cos^2(\phi) ,\\ \mu^2(\delta) &=& \mu^2(\theta_M)\cos^2(\phi) + \mu^2(\theta_m)\sin^2(\phi) . \end{eqnarray} \subsubsection{Calibration errors} In the bright source regime, the error on the flux density is dominated not by noise but by calibration errors. This is added in quadrature to the other sources of error as a fractional error of 4\% on the integrated flux density. This was determined by measuring the local increase in RMS flux density around bright sources of known flux density, where calibration errors are the dominant source of noise, and dividing it by the flux density of each bright source. \subsection{Remaining systematic errors} \subsubsection{Error on the absolute calibration} \label{sect:abserr} \begin{figure*}[p] \centering \includegraphics[width=0.8\textwidth]{flux_ratio_nogrid.png} \caption{Plotted against Dec, the ratio of 180\,MHz flux density as predicted from catalogued values to that determined in our survey, for cross-matched unresolved sources. Unfilled diamonds represent 210 unresolved Culgoora sources with the 160\,MHz flux density scaled to a 180\,MHz flux density using the MWACS spectral index. Greyscale is the MWACS source flux density in Jy. Other points are fits to sources in the MRC and MS4 samples (as described in Section~\ref{sect:abserr}). Dots represent the 32~sources fit best by a curved spectrum. Crosses show 21~sources without curved spectra whose 180\,MHz flux was extrapolated. Circles show the 32~sources classified as neither curved nor extrapolated.} \label{fig:bh} \end{figure*} \begin{figure*}[p] \centering \begin{subfigure}[b]{0.49\textwidth} \centering \includegraphics[width=\textwidth]{cumulative_ratio_linear.png} \end{subfigure} \begin{subfigure}[b]{0.49\textwidth} \centering \includegraphics[width=\textwidth]{cumulative_ratio_log.png} \end{subfigure} \caption{The cumulative histograms of the difference of the flux density ratio from unity, where the ratio is the ratio of the flux densities of unresolved sources found in MWACS to the flux densities as measured by other catalogues (see Section~\ref{sect:abserr}). The grey line shows Culgoora, the thin black line shows all MS4 and MRC4 sources, and the thick black line shows only those MS4 and MRC4 sources which do not have curved spectra and are not pure extrapolations from 408\,MHz downward. The left panel shows the histogram in linear intervals, while the right panel shows the histogram with a log scale.\label{fig:flux_ratio_histogram}} \end{figure*} To test the consistency of the absolute flux density calibration across the field and across different sources, we first crossmatched our catalogue with the only other catalogue which lies within our frequency range: the Culgoora-3 list of radio source measurements at 160\,MHz, at a resolution of 1\farcm85. These flux densities were scaled to 180\,MHz using the MWACS spectral index and the scaled flux densities compared with the MWACS flux densities, shown in Figure~\ref{fig:bh}. While the scatter in the ratio of flux densities is large, and considerably larger than the estimated error on the Culgoora fluxes \citep[][Tables~1\&2]{1977AuJPA..43....1S}, there is no evidence of a systematic offset with respect to Declination, which would have indicated uncorrected primary beam errors. In addition, we identified all 41 sources which appeared in all the five catalogues used for absolute calibration (Section~\ref{sect:abscorrect}) and which had an MRC (408\,MHz) flux density greater than 4\,Jy (hereafter the MRC4 sample). Since these catalogues only cover a restricted Dec range near the top of our field, these sources were supplemented by 73 sources from the ``MS4'' sample, defined by \citet{2006AJ....131..100B} as those sources in the MRC catalogue with a flux density greater than 4\,Jy.\footnote{In Section 4 of \citet{2006AJ....131..100B}, the authors present extrapolated 178\,MHz flux densities; we do not make use of these data.} From the resulting sample we excluded all sources which were resolved in our catalogue -- defined as those with a fitted ellipse $>5$\% larger than synthesised beam ellipse (19/73 MRC4 sources and 9/41 MS4 sources were excluded). Two fits were made to the catalogued flux densities: one a simple power law (i.e. a straight-line fit in log-log space); another incorporating curvature (a parabolic fit in log-log space). Those sources where the 180\,MHz flux density predicted by the two fits were discrepant by more than 10\% were classified as ``curved''. For the MS4 sources we used the multi-frequency flux density measurements compiled by \citet{2006AJ....131..100B}. Many of these sources had no listed flux density below 300\,MHz (i.e. the 180\,MHz flux density is extrapolated from flux density measurements at 365 or 408\,MHz and above). These sources were classified as ``extrapolated''. The resulting flux density ratios are also shown in Figure~\ref{fig:bh}, which overall illustrates the challenge of defining a low-frequency flux scale in the Southern Hemisphere. There are few sources with well-characterised spectra at this frequency, and there are inconsistencies between these sources. Figure~\ref{fig:flux_ratio_histogram} replots these data as cumulative histograms of the difference of the flux density ratio from unity. Three subsets of (unresolved) sources are shown: Culgoora, MS4 \& MRC4, and MS4 \& MRC4 sources which do not have curved spectra and are not pure extrapolations from 408\,MHz downward. Using this latter, most high-quality subset, we see that 68\% of sources lie within 10\% of unity. We therefore add an absolute flux calibration error of 10\% in quadrature to the final flux density error after fitting. \subsubsection{Primary-beam errors} For a mosaicked drift scan, an error in the primary-beam model would be expected to manifest itself as a Dec-dependent error in flux density. Overall there is little evidence of a Dec-dependent error in Figure~\ref{fig:bh}. Any error in the primary beam model would very likely have strong frequency-dependence, such that it may systematically affect the spectral index. This can be tested by binning the sources by spectral index and looking for systematic changes with Dec. The results of this analysis are shown in Figure~\ref{fig:alpha_de}. \begin{figure} \centering \includegraphics[width=\columnwidth]{dec_alpha_quartiles.png} \caption{Quartiles of spectral index (filled circles indicate median, bars indicate upper and lower quartiles) of sources in each of 20 equal-width bins covering the full range of Dec. The shaded region indicates the declination range of the overlap region; the dotted vertical line indicates the declination of 3C32.} \label{fig:alpha_de} \end{figure} There are clearly systematic changes in the median spectral index with Dec, though there is no obvious functional form which could be modelled to take account of this. However the change in spectral index across the band corresponds to an error across the frequency range of less than 10\%. An error of 0.1 is added in quadrature to the to the error on the fitted spectral index to take account of the systematic shift in spectral index seen here. \subsubsection{Astrometry errors}\label{sec:astrometry} Figure~\ref{fig:astrometry} shows the astrometric error for all unresolved sources cross-matched between our catalogue and the NVSS catalogue for the Dec\,$-27$ mosaic, and the SUMSS catalogue for the Dec\,$-47$ mosaic. The astrometric errors in RA and Dec are not entirely correlated, vary with RA (time, for a drift scan), and are likely to be functions of ionospheric activity and viewing angle through the ionosphere (i.e. zenith angle). As we do not attempte to solve for this systematic error, we measure the RMS without weighting each measurement by the flux density of the sources, which leads to a conservative astrometric RMS estimate of 0\farcm3. This value is added in quadrature to the errors on the RA and Dec derived in Section~\ref{sect:noise-like-errors}. \begin{figure}[h!] \includegraphics[width=0.5\textwidth]{astrometry.png} \caption{Astrometric offsets (left: RA; right: Dec) for the 180\,MHz mosaics, shown against RA (effectively, time of night). Top: Dec\,$-27$ mosaic, cross-matched against NVSS. Bottom: Dec\,$-47$ mosaic, cross-matched against SUMSS. The colour axis is the log$_{10}$ of the source flux density in Jy (as measured in the 180\,MHz mosaic before amplitude calibration).} \label{fig:astrometry} \end{figure} \subsection{Resolved sources}\label{sect:multi-comps} While the majority of sources in the catalogue are unresolved at the resolution of the commissioning array, and are therefore described well by a single flux density, some objects are more complex and must be examined more closely to determine an integrated flux density. These are automatically detected as described in Section~\ref{sect:goodbadsplit}. For each of these sources, a $100\times100$ pixel subsection of the mosaic was regridded to a local SIN projection, and the appropriate point spread function determined in Section~\ref{subsect:imaging} was added to the header. The source-finding software \textsc{Aegean} was then run on the sub-image, detecting and fitting multiple components simultaneously. The results of this fitting at the highest frequency only were used to produce the integrated flux density measurement at 180\,MHz. Due to the varying resolution over the bandwidth of the array, cross-matching from the 180\,MHz-detected sources to the other two frequencies is not always straight-forward when sources are extended, or close to each other. \begin{itemize} \item{For some sources, particularly those which are only slightly extended and not confused with other nearby sources, the components match easily, and integrated flux densities are calculated for each freqency.} \item{For confused sources, where different numbers of components are detected at different frequencies, an integrated flux density is calculated for each frequency. This is done by flood-filling an \textsc{Aegean}-detected island down to a level four times the local RMS flux density, and integrating.} \item{For those ($\approx5\sigma$) sources which were not easily fit by a Gaussian, or by a flood-fill, at 120 and 150\,MHz, the position centroid from the source fit to the 180\,MHz data was used to perform a `forced' measurement on that pixel position in the lower-frequency data.} \end{itemize} The flux densities at each frequency are then fit in the usual way to produce a single spectral index for the 180-MHz source components. Examples of each method are shown in Figure~\ref{fig:extended_sources}. The fitted spectral indices and 180\,MHz flux densities for the components of resolved sources are reported in the catalogue in the same way as for the unresolved sources. As shown in Table~\ref{tab:catalogue-extract}, a column in the table indicates whether the spectral index was determined by integrating an extended source, or performing a forced measurement. Error measurements for the flux densities of these extended sources were identical to those discussed in Section~\ref{sect:errors}, with the fitting error combined in quadrature with the RMS noise. It should be noted, however, that the changing resolution of the instrument at different frequencies, combined with the intrinsic complexity of these sources, mean that the automatically derived spectral indices should be treated with caution. Postage stamps of the extended sources are available on request should the reader wish to perform their own measurements. \begin{figure*}[p] \centering \begin{subfigure}[b]{0.90\textwidth} \centering \includegraphics[angle=0,width=0.90\textwidth]{crossmatched-measurement.png} \caption{\label{fig:cross-match-extended}} \end{subfigure} \begin{subfigure}[b]{0.90\textwidth} \centering \includegraphics[angle=0,width=0.90\textwidth]{island-measurement.png} \caption{\label{fig:island-flux}} \end{subfigure} \begin{subfigure}[b]{0.90\textwidth} \centering \includegraphics[angle=0,width=0.90\textwidth]{forced-measurement.png} \caption{\label{fig:forced-measurement}} \end{subfigure} \caption{Examples of measuring extended emission; top: cross-matching extended components; middle: integrating a flux density where the source is comprised of multiple components; bottom: forcing measurements on faint sources. In each case, the 180\,MHz image is shown on the left, the 150\,MHz image in the middle, and the 119\,MHz image on the right. All images are square and measure 21\hbox{$^\prime$\,} on each side. The colourscales are linear and the same for each source; the minimum is always $-0.2$Jy\,beam$^{-1}$ and the maxima are 3, 1.5 and 0.4\,Jy\,beam$^{-1}$, from top to bottom. The white contours are linear at $5\sigma$ levels, starting at $4\sigma$; left-to-right, top-to-bottom, the local RMS $\sigma=100$, 60, 42, 71, 47, 30, 54, 30, and 23\,mJy\,beam$^{-1}$. The effective PSF is shown as a filled ellipse in the bottom-left corner. (a) The source is detected as a single extended object at all frequencies, so the \textsc{Aegean} component fits are cross-matched, and the integrated flux densities fitted to produce $S_\mathrm{180\,MHz}$ and $\alpha$ as for unresolved sources. (b) For multiple-component sources where \textsc{Aegean} was unable to detect separate components at the lower frequencies, the components detected at 180\,MHz are used to set up peaks from which the 150\, and 119\,MHz images are flood-filled down to a level of $4\sigma$, shown as a thicker contour. The integrated flux densities of the measurements at all three frequencies are used to fit a spectral index for all of the components resolved and reported at 180\,MHz. (c) For those sources which are not well-fit by a Gaussian, and are very close to $5\sigma$ at the lower frequencies, we force measurements at the position (indicated by a cross) of the 180\,MHz centroid (black ellipse) on the 150 and 120\,MHz mosaics. All three flux density measurements are used to fit a spectral index for the components resolved and reported at 180\,MHz.} \label{fig:extended_sources} \end{figure*} \section{Source catalogue} \label{sec:sourcecat} The MWACS catalogue comprises {14,110} sources and is available via Vizier\footnote{http://vizier.u-strasbg.fr/}, and in the electronic edition of this journal. We report the positions of sources, the 180\,MHz integrated flux density and spectral index, the Gaussian fit parameters (major axis, minor axis, and position angle) as well as errors on all of the above parameters. Flags indicate whether a source was confused or extended and thus required special treatment (see Section~\ref{sect:multi-comps}.) The estimated synthesised beam at that position is also provided. Postage stamps are available on request. \begin{sidewaystable*}[p] \footnotesize \centering \caption{Source Catalogue (only the first 15 sources are shown) Table columns are defined as follows: (1) IAU name hhmm.m+ddmm. (2) RA. (3) Error on RA. (4) Dec. (5) Error on Dec. (6) Integrated Flux density at 180\,MHz. (7) Error on Integrated flux density. (8) Spectral index ($\alpha$ where $S_{\nu}\propto\nu^{\alpha}$). (9) Error on spectral index. (10) Major axis of source. (11) Error on major axis. (12) Minor axis of source. (13) Error on minor axis. (14) Position angle of source. (15) Error on position angle of source. (NB 10, 12 \& 14, are as measured in the 180\,MHz maps and include convolution with the synthesised beam). (16, 17, 18) Major axis, minor axis and position angle of the PSF at location of source. (19) Source index within Dec\,$-27$ field. (20) Source index within Dec\,$-47$ field. (21) For sources where a multi-component fit was required: the index of the component (NB sources in the overlap region are identified by having non-null values for both 19\&20. 19, 20 \& 21 can be used to cross-match multiple components. All components will share a value for columns 19 and/or 20 with each component source having a unique value for 21). (22) Type of spectral fit used to determine source spectral index; 0 indicates source fitting at all three frequencies (Figure~\ref{fig:cross-match-extended}); 1 indicates floodfill (Figure~\ref{fig:island-flux}); 2 indicates a forced measurement (Figure~\ref{fig:forced-measurement}). }\label{tab:catalogue-extract} \begin{tabular}{lrrrrrrrrrr} \hline \multicolumn{1}{c}{Name} & \multicolumn{1}{c}{RAJ2000} & \multicolumn{1}{c}{e\_RAJ2000} & \multicolumn{1}{c}{DEJ2000} & \multicolumn{1}{c}{e\_DEJ2000} & \multicolumn{1}{c}{S180} & \multicolumn{1}{c}{e\_S180} & \multicolumn{1}{c}{SpIndex} & \multicolumn{1}{c}{e\_SpIndex} & \multicolumn{1}{c}{MajAxis} & \multicolumn{1}{c}{e\_MajAxis} \\ \multicolumn{1}{c}{} & \multicolumn{1}{c}{deg} & \multicolumn{1}{c}{arcmin} & \multicolumn{1}{c}{deg} & \multicolumn{1}{c}{arcmin} & \multicolumn{1}{c}{Jy} & \multicolumn{1}{c}{Jy} & \multicolumn{1}{c}{} & \multicolumn{1}{c}{} & \multicolumn{1}{c}{arcmin} & \multicolumn{1}{c}{arcmin} \\ \multicolumn{1}{c}{(1)} & \multicolumn{1}{c}{(2)} & \multicolumn{1}{c}{(3)} & \multicolumn{1}{c}{(4)} & \multicolumn{1}{c}{(5)} & \multicolumn{1}{c}{(6)} & \multicolumn{1}{c}{(7)} & \multicolumn{1}{c}{(8)} & \multicolumn{1}{c}{(9)} & \multicolumn{1}{c}{(10)} & \multicolumn{1}{c}{(11)} \\ \hline MWACSJ0000.0-1704 &0.003 &0.5 &-17.070 &0.4 &0.66 &0.09 &-1.0 &0.4 &4.5 &0.4 \\ MWACSJ0000.1-2824 &0.016 &0.5 &-28.406 &0.5 &0.32 &0.04 &-1.1 &0.4 &4.2 &0.4 \\ MWACSJ0000.1-4617 &0.028 &0.6 &-46.297 &0.5 &0.50 &0.07 &-0.8 &0.4 &5.4 &0.5 \\ MWACSJ0000.1-4910 &0.031 &0.4 &-49.182 &0.4 &0.78 &0.09 &-1.5 &0.2 &4.3 &0.2 \\ MWACSJ0000.1-5234 &0.032 &0.6 &-52.577 &0.6 &0.21 &0.04 &-1.7 &0.6 &4.2 &0.6 \\ MWACSJ0000.2-4333 &0.042 &0.5 &-43.553 &0.5 &0.30 &0.04 &-1.4 &0.4 &4.3 &0.5 \\ MWACSJ0000.3-2450 &0.080 &0.7 &-24.844 &0.5 &0.27 &0.04 &-0.4 &0.7 &5.0 &0.7 \\ MWACSJ0000.3-2724 &0.076 &0.6 &-27.414 &0.5 &0.27 &0.04 &-6.1 &0.3 &4.8 &0.6 \\ MWACSJ0000.3-3410 &0.067 &0.5 &-34.174 &0.4 &0.63 &0.08 &-1.1 &0.3 &4.6 &0.4 \\ MWACSJ0000.4-3024 &0.091 &0.7 &-30.409 &0.6 &0.20 &0.04 &-1.2 &0.8 &4.0 &0.6 \\ MWACSJ0000.4-3822 &0.107 &0.4 &-38.383 &0.4 &0.72 &0.09 &-0.7 &0.3 &4.3 &0.3 \\ MWACSJ0000.4-4721 &0.105 &0.4 &-47.363 &0.3 &1.11 &0.12 &-0.7 &0.2 &4.3 &0.2 \\ MWACSJ0000.4-5635 &0.107 &0.5 &-56.589 &0.5 &0.53 &0.07 &-1.4 &0.4 &4.7 &0.5 \\ MWACSJ0000.5-3320 &0.119 &0.5 &-33.337 &0.4 &0.68 &0.09 &-0.6 &0.3 &4.7 &0.4 \\ MWACSJ0000.5-3452 &0.125 &0.5 &-34.875 &0.4 &0.90 &0.11 &-1.4 &0.3 &4.5 &0.4 \\ \hline\end{tabular} \begin{tabular}{rrrrrrrrrcr} \hline \multicolumn{1}{c}{MinAxis} & \multicolumn{1}{c}{e\_MinAxis} & \multicolumn{1}{c}{PA} & \multicolumn{1}{c}{e\_PA} & \multicolumn{1}{c}{MajAxisBeam} & \multicolumn{1}{c}{MinAxisBeam} & \multicolumn{1}{c}{PABeam} & \multicolumn{1}{c}{ID\_C102} & \multicolumn{1}{c}{ID\_C103} & \multicolumn{1}{c}{Component\_ID} & \multicolumn{1}{c}{Fit} \\ \multicolumn{1}{c}{arcmin} & \multicolumn{1}{c}{arcmin} & \multicolumn{1}{c}{deg} & \multicolumn{1}{c}{deg} & \multicolumn{1}{c}{arcmin} & \multicolumn{1}{c}{arcmin} & \multicolumn{1}{c}{deg} & \multicolumn{1}{c}{} & \multicolumn{1}{c}{} & \multicolumn{1}{c}{} & \multicolumn{1}{c}{} \\ \multicolumn{1}{c}{(12)} & \multicolumn{1}{c}{(13)} & \multicolumn{1}{c}{(14)} & \multicolumn{1}{c}{(15)} & \multicolumn{1}{c}{(16)} & \multicolumn{1}{c}{(17)} & \multicolumn{1}{c}{(18)} & \multicolumn{1}{c}{(19)} & \multicolumn{1}{c}{(20)} & \multicolumn{1}{c}{(21)} & \multicolumn{1}{c}{(22)} \\ \hline 2.9 &0.3 &-67 &27 &4.4 &2.8 &-63 &4760 &\dots & &0 \\ 3.0 &0.3 &-56 &32 &4.4 &2.8 &-64 &4587 &\dots & &0 \\ 3.3 &0.3 &-60 &25 &4.3 &2.9 &-51 &\dots &4229 &\dots &0 \\ 3.1 &0.2 &-45 &24 &4.3 &3.0 &-50 &\dots &1620 &\dots &0 \\ 3.0 &0.4 &-53 &36 &4.4 &3.0 &-48 &\dots &6683 &\dots &0 \\ 3.1 &0.3 &-59 &32 &4.3 &2.9 &-52 &\dots &4870 &\dots &0 \\ 2.9 &0.4 &-74 &29 &4.4 &2.8 &-64 &7128 &\dots & &0 \\ 2.7 &0.3 &-56 &26 &4.4 &2.8 &-64 &5995 &\dots &1 &1 \\ 3.0 &0.3 &-66 &26 &4.4 &2.8 &-63 &4910 &\dots & &0 \\ 2.8 &0.4 &-67 &38 &4.4 &2.8 &-63 &6951 &\dots & &0 \\ 3.0 &0.2 &-57 &26 &4.2 &2.9 &-54 &\dots &3425 &\dots &0 \\ 3.0 &0.2 &-53 &21 &4.3 &2.9 &-50 &\dots &1358 &\dots &0 \\ 3.1 &0.3 &-41 &28 &4.5 &3.1 &-45 &\dots &4385 &\dots &0 \\ 3.0 &0.3 &-64 &24 &4.4 &2.8 &-63 &3834 &\dots & &0 \\ 2.9 &0.2 &-68 &24 &4.4 &2.8 &-63 &3443 &\dots & &0 \\ \hline\end{tabular} \end{sidewaystable*} \subsection{Catalogue reliability and completeness} Recall that all of the source-finding for this survey was performed on the highest-frequency mosaic. For point sources, after source identification, the source was fitted at all other frequencies. If any of these fits failed, the source would be treated as resolved; see Section~\ref{sect:multi-comps} for details. We searched for and identified ten spurious sources in the sidelobes of bright sources by searching the areas around ($\approx100$) sources of $S>10$\,Jy, comparing images at all frequencies, and removed these from the catalogue; the faintest source which produced a spurious sidelobe source was 14\,Jy. In typical circumstances away from bright sources, we require that all sources are detectable at $5\sigma$ at 180\,MHz and well-fit by a Gaussian in the other bands, so with a Gaussian random background, the chance of any source being spurious is vanishingly small. Our noise is somewhat non-Gaussian due to the presence of undeconvolved source sidelobes, but in the general field, these sidelobes are faint due to our good $u,v$-coverage, which suppresses sidelobes down to a $10\%$ level even before deconvolution; and the positions of change sidelobe alignments would vary with frequency and thus be unlikely to be detected in every channel. Therefore we can have an extremely high confidence that all of the sources in the catalogue are real. The varying detection threshold affects the catalogue completeness for two reasons. More obviously, the variations in the local RMS cause a varying absolute detection limit across the field. In addition, the noise in the map introduces a non-zero probability that a source whose intrinsic flux is $>5\sigma$, has a measured flux $<5\sigma$ due to instrumental effects (and vice versa). We therefore present the noise properties of our maps in some detail. \subsubsection{Variations in RMS sensitivity}\label{sect:noise} Figure~\ref{fig:rmsmap} shows maps of the local RMS of the mosaics rescaled to take into account absolute calibration (with XX and YY combined together for the Dec\,$-47$ scan). This information is summarised in Figure~\ref{fig:rms_cdf}. \begin{figure} \centering \includegraphics[width=0.5\textwidth]{rms_cumulative.png} \caption{Cumulative distribution functions of the RMS at 180\,MHz for those areas of the map outside of the overlap region for the Dec\,$-27$ scan (black dashed line) and Dec\,$-47$ scan (black dotted line). A subset of data truncated in RA and Dec range is also shown for both scans: Dec\,$-27$ (grey dashed line: $22^{\mathrm{h}}00^{\mathrm{m}}<\alpha<07^{\mathrm{h}}30^{\mathrm{m}}$, $-30\arcdeg<\delta<-20\arcdeg$); Dec\,$-47$ (grey dotted line: $21^{\mathrm{h}}15^{\mathrm{m}}<\alpha<06^{\mathrm{h}}40^{\mathrm{m}}$, $-42\arcdeg<\delta<-52\arcdeg$)} \label{fig:rms_cdf} \end{figure} \begin{figure \centering \begin{subfigure}{0.5\textwidth} \centering \includegraphics[width=\textwidth]{histogram_c102.png} \label{fig:c102_hist} \end{subfigure} \begin{subfigure}{0.5\textwidth} \centering \includegraphics[width=\textwidth]{histogram_c103.png} \label{fig:c103_hist} \end{subfigure} \caption{Histogram of the pixel brightnesses of the 180\,MHz maps (P(D)). RA and Dec ranges are $22^{\mathrm{h}}00^{\mathrm{m}}<\alpha<07^{\mathrm{h}}30^{\mathrm{m}}$, $-30\arcdeg<\delta<-20\arcdeg$ (approx 1300 square degrees) for the Dec\,$-27$ mosaic (top) and $21^{\mathrm{h}}15^{\mathrm{m}}<\alpha<06^{\mathrm{h}}40^{\mathrm{m}}$, $-42\arcdeg<\delta<-52\arcdeg$ (approx 1000 square degrees) for the Dec\,$-47$ mosaic (bottom). The RMS for each pixel distribution (determined via the semi-interhexile range) is 0.035\,Jy\,beam$^{-1}$ and 0.029\,Jy\,beam$^{-1}$ respectively. The dashed line is a gaussian curve with this RMS, centred on the median flux density ($-2.4$\,mJy\,beam$^{-1}$ and $-3.2$\,mJy\,beam$^{-1}$ respectively).} \label{fig:hist} \end{figure} This figure shows that the extremely high RMS regions associated with bright extended sources, particularly near the Galactic plane, account for a small proportion of survey area. Overall, the sensitivity of the survey varies about the median of $\sim$40\,mJy by less than a factor of two over 80\% of the survey area, a variation in sensitivity comparable with the VLSS \citep[c.f.][Figure~7]{2007AJ....134.1245C}. The RMS in the Dec\,$-27$ and Dec\,$-47$ mosaics are reasonably comparable, though the Dec\,$-47$ mosaic has a somewhat lower RMS, particularly for regions below the median. We ascribe this to a better phase calibration for the Dec\,$-47$ data, as a brighter phase calibrator was used. The RMS becomes only marginally worse towards the edge of the primary beam in the Dec\,$-47$ mosaic, however it becomes significantly worse towards the edge of the primary beam in the Dec\,$-27$ mosaic. This is because the sky rotates more quickly through the primary beam closer to the celestial equator. Figure~\ref{fig:hist} shows the histogram of flux-density values across a significant proportion of each mosaic. When calculated over a wide area, the RMS for the two maps agrees to $\sim$10\%; 90\% of the surveyed area has an RMS lower than 50\,mJy, except for the very Northern edge. \subsection{Fornax~A} Fornax~A is a nearby ($z\approx0.006$) radio galaxy, situated in the Fornax cluster of galaxies. It comprises two large lobes extending NNW and SSE 20\hbox{$^\prime$\,} ($\approx200$\,kpc) from the lenticular galaxy NGC~1316, at RA$=03^{\mathrm{h}}22^{\mathrm{m}}41.789^{\mathrm{s}}$ Dec$=-37^{\circ}12\arcmin29\farcs52$. On scales $\gtrsim20$\hbox{$^\prime$\,}, structure is not sampled by MWACS, so we cannot measure the integrated flux density for this object, only set lower limits. This under-sampling also causes deconvolution errors across our mosaics, resulting in increased noise levels around the Fornax~A region. \begin{figure} \centering \includegraphics[scale=1, angle=-90, width=0.5\textwidth]{FornaxA.png} \caption{Fornax~A at 180\,MHz; contours are linear at 10\% levels from the peak brightness of 16.75\,Jy\,beam$^{-1}$. The synthesised beam is shown as an open ellipse at the bottom-left of the image, with dimensions $4\farcm22\times2\farcm85$, position angle $-53\fdg8$. The RMS of this image is 92\,mJy.}\label{fig:fornax} \end{figure} Using \textsc{BlobCat} to flood-fill the entire region down to a $4\sigma$ level and measure the integrated flux density, we set lower limits across the entirety of Fornax~A at $786\pm 39$, $668\pm 33$, and $514\pm 26$\,Jy at 120, 150 and 180\,MHz, respectively; these are consistent with the values measured by \citet{2013ApJ...771..105B} and \citet{2014McKinley}. An image of the radio galaxy at the highest MWACS resolution is shown in Figure~\ref{fig:fornax}. \section{Discussion and conclusions} \label{sec:conclusions} We have performed a survey of approximately {6,100} square degrees of the southern sky during the MWA's commissioning period in 2012~October: the MWA Commissioning Survey (MWACS). The observations were made in meridian drift scan mode at two Dec settings ($-26\fdg7$ and $-47\fdg5$) centred on three frequencies: 119, 150 and 180\,MHz, resulting in an effective frequency coverage of 104 to 196 MHz. Because the data were taken as drift scans, and because the MWA's primary beam X and Y polarisation responses are different when off-zenith, some unconventional data processing was required. Six separate mosaic images were formed, one for each frequency and Dec covering the entire RA range of the drift scan. By taking advantage of the instantaneous (near) co-planarity of the array and good instantaneous $u,v$-coverage (including large fractional bandwidth), the data were processed as continuum multi-frequency synthesis images (at the three central frequencies) by forming a mosaic of weighted regridded snapshot images. The source 3C32 was used to set the absolute flux density scale. Stokes~I images at 119, 150 and 180\,MHz were formed by adding the primary-beam weighted XX and YY mosaics. The absolute flux density was bootstrapped from the Dec\,$-27$ mosaic to the Dec\,$-47$ mosaic using sources in the overlapping five degrees of Declination. We found the primary beam response at Dec\,$-47$ to be sufficiently different between the XX and YY polarisations that an independent scaling factor was used for each before being combined to Stokes~I. The data do not have sufficient short baselines to recover large-scale structure, so only compact sources were extracted from the mosaics. We first identified sources as islands of $>5\sigma$ flux density, then separated sources into single-component (unresolved) and multi-component (resolved, and/or multiple nearby components) categories, and measured their flux densities at each frequency, appropriately. We present the result of the MWACS a single catalogue comprised of 180-MHz flux densities and spectral indices calculated from the three individual measured frequencies, for {14,110} sources. The sensitivity of MWACS, as measured by local image RMS, is mostly consistent around 40\,mJy\,beam$^{-1}$ ($1\sigma$) across the mosaics but degrades near bright sources and near the edges of the fields. All sources found in MWACS at 180\,MHz were also found in the two lower frequency mosaics and had reasonable spectra given their S/N. We therefore consider detections to be reliable subject to the usual considerations on confusion and resolution. Users are advised, however, that non-detections can only be used to set an upper limit to flux density in the context of the local RMS. Regions within a few degrees of the edge of the image or bright sources will have locally higher RMS. This commissioning survey demonstrates the impressive survey capabilities of the MWA and the feasibility of processing MWA drift scan datasets using fairly conventional radio astronomy tools with simple mosaicking techniques. The full MWA has near-complete snapshot $u,v$-coverage between approximately 5~and~500 wavelengths (as multi-frequency synthesis images) and with 128 fully cross-correlated tiles, its quality of calibration is higher than our combined commissioning sub-arrays. We therefore expect the GLEAM survey to be confusion limited in Stokes~I, to be able to capture large-scale structures (up to $\sim$10\arcdeg) and to have much reduced imaging artifacts from better calibration and deconvolution. The progression of the MWA from 32T prototype (circa 2009--2011) through the commissioning array (late 2012 through mid-2013) and into operations (mid-2013 on) has shown a steady improvement in area and depth of sky coverage and of our understanding of the instrument and data. The GLEAM survey, currently underway, has derived great benefit from the accumulated expertise of the 32T and commissioning phases and will likewise provide increased understanding of the instrument, especially the primary beam and resulting polarisation performance of the MWA, and data processing methods. Looking forward to SKA-low, we expect the expertise gained from processing the large GLEAM dataset on world-class supercomputers to be provide equally valuable lessons when SKA-low Phase\,1 is complete. \section*{Acknowledgements} This scientific work makes use of the Murchison Radio-astronomy Observatory, operated by CSIRO. We acknowledge the Wajarri Yamatji people as the traditional owners of the Observatory site. Support for the MWA comes from the U.S. National Science Foundation (grants AST-0457585, PHY-0835713, CAREER-0847753, and AST-0908884), the Australian Research Council (LIEF grants LE0775621 and LE0882938), the U.S. Air Force Office of Scientific Research (grant FA9550-0510247), and the Centre for All-sky Astrophysics (an Australian Research Council Centre of Excellence funded by grant CE110001020). Support is also provided by the Smithsonian Astrophysical Observatory, the MIT School of Science, the Raman Research Institute, the Australian National University, and the Victoria University of Wellington (via grant MED-E1799 from the New Zealand Ministry of Economic Development and an IBM Shared University Research Grant). The Australian Federal government provides additional support via the Commonwealth Scientific and Industrial Research Organisation (CSIRO), National Collaborative Research Infrastructure Strategy, Education Investment Fund, and the Australia India Strategic Research Fund, and Astronomy Australia Limited, under contract to Curtin University. We acknowledge the iVEC Petabyte Data Store, the Initiative in Innovative Computing and the CUDA Center for Excellence sponsored by NVIDIA at Harvard University, and the International Centre for Radio Astronomy Research (ICRAR), a Joint Venture of Curtin University and The University of Western Australia, funded by the Western Australian State government. This research has made use of the NASA/IPAC Extragalactic Database (NED) which is operated by the Jet Propulsion Laboratory, California Institute of Technology, under contract with the National Aeronautics and Space Administration. \newcommand{PASA}{PASA} \newcommand{AJ}{AJ} \newcommand{\apj}{ApJ \newcommand{ApJS}{ApJS} \newcommand{ApJL}{ApJL} \newcommand{A{\&}A}{A{\&}A} \newcommand{A{\&}AS}{A{\&}AS} \newcommand{MNRAS}{MNRAS} \newcommand{ARAA}{ARAA} \newcommand{PASP}{PASP} \bibliographystyle{apj}
\section{Introduction} Nonequilibrium phase transitions into absorbing states describe several problems, such as wetting phenomena, spreading of diseases, chemical reactions and others \cite{r1,r2}. In the last years, a large effort for its characterization including experimental verifications \cite{r3,r3.1,r3.2} and the establishment of distinct universality classes \cite{r1,r2,chate} have been undertaken. Generically, continuous phase transitions into an absorbing state for systems without conservation laws nor extra symmetries fall into in the directed percolation (DP) universality class \cite{r2,r4,r5}. The best example of this category is probably the contact process (CP) \cite{r4}. It is defined on a given $d$-dimensional lattice and the dynamics comprehends the creation in the presence of at least one adjacent particle and spontaneous annihilation. Since the particle creation requires the presence of adjacent particles, a configuration devoided of species is absorbing. Conversely, discontinuous absorbing transitions also appear in different systems \cite{zgb,bid,r7,grass,r9}, but comparatively they have received less attention than the continuous ones. Recently, they have attracted interest for the search of minimal ingredients for their occurrence, being one possibility the so called ``restrictive'' (threshold) contact processes (CPs). These models are variants of the second Sch\"ogl model, in which the particle creation is similar to the usual CP, but instead it requires a minimal neighborhood larger than 1 particle for creating a new species \cite{evans,evans2,oliveira2,oliveira,carlos}. Different studies have stated that the phase transition, induced by this mild change (with respect to the usual CP), remains unaffected under the inclusion of distinct creation \cite{evans,evans2,oliveira2,oliveira,carlos} and annihilation rules \cite{carlos} as well as for different lattice topologies \cite{durret}. In some specific cases \cite{evans,evans2}, in which the particle creation occurs only if one has at least two adjacent diagonal pairs of particles, the phase transition is characterized by a generic two-phase coexistence and exhibits an interface orientational dependence at the transition point. On the other hand, when the transition rates depend only on the number of nearest neighbors particles (and not their orientations) the discontinuous transitions take place at a single point \cite{oliveira2, oliveira,carlos}. Despite the apparent robustness of first-order transitions for the above mentioned restrictive examples, the effect of some (relevant) dynamics has been so far unexplored in the present context. These dynamics, such as spatial disorder and particle diffusion, can cause drastic changes in continuous phase transitions \cite{r2}. One of the few available studies shows that spatial disorder suppresses the phase coexistence, giving rise to a continuous transition belonging to a new universality class \cite{munoz14}. A similar scenario of scarce results also holds for the outcome of diffusion. Very recently, a stochastic differential equation (such as a Langevin equation) reported that different diffusion strengths lead to opposite findings (in two dimensions). Whenever the transition is discontinuous for larger diffusion values, limited rates suppress it, giving rise to a critical phase transition belonging to the DP universality class \cite{munoz214}. With these ideas in mind, we give a further step by tackling the influence of diffusion in lattice systems presenting discontinuous absorbing phase transitions. \cite{oliveira2,oliveira}. Our study aims to answer three fundamental points: (i) what is the effect of strong and limited diffusion in these cases? (ii) does it suppress the discontinuous transition? (iii) How does our results compare with those obtained from the coarse-grained description in Ref. \cite{munoz214}? In other words, are there differences between lattice models and Langevin equations? We consider two lattice models and three representative values of diffusion, in order to exemplify the low, intermediate and large regimes. Models will be studied via MFT and distinct kinds of numerical simulations (explained further). Results suggest that the discontinuous transition is maintained in both models for all diffusion values. Also, a finite-size scaling behavior similar to discontinuous transitions studied in Ref. \cite{carlos} has been found. Since there is no theory for the nonequilibrium case, our results can shed light over a general finite-size scaling for them.\\ This paper is organized as follows: Sec. II presents the models and mean-field analysis. Sec. III shows the numerical results and finally conclusions are drawn in Sec. IV. \section{Models and mean-field analysis} Let us consider systems of interacting particles placed on a square lattice of linear size $L$. Each site has an occupation variable $\eta_i$ that assumes the value $0\,\,(1)$ whenever sites are empty (occupied). The model A is defined by the following interaction rules: particles are annihilated with rate $\alpha$ and are created in empty sites only if their number of nearest neighbor particles $nn$ is larger than 1 ($nn \ge 2$), with rate $nn/4$ \cite{oliveira2}. There is no particle creation if $nn \le 1$. The model B is similar to the model A, but the particle creation rate always reads $1$, provided $nn \ge 2$ \cite{oliveira}. Thus, whenever particles are created with rate proportional to the number of their nearest neighbors for the model A, it is independent on $nn$ in the model B. Besides the above creation-annihilation dynamics, each particle also hops to one of its nearest neighbor sites with probability $D$, provided it is empty. In the regime of low annihilation parameters, the system exhibits indefinite activity in which particles are continuously created and destroyed. In contrast, for larger $\alpha$'s, the system is constrained in the absorbing phase. The phase transition separates above regimes at a transition point $\alpha_{0}$. In the absence of diffusion, the phase transitions for both models A and B are discontinuous and occur at $\alpha_0=0.2007(6)$ and $0.352(1)$, respectively \cite{oliveira2,oliveira}. The first inspection of the effect of diffusion can be achieved by performing mean-field analysis. The starting point is to write down the time evolution of relevant quantities from the interaction rules and truncating the associated probabilities at a given level. Since the diffusion conserves the number of particles, it is required to take into account at least correlations of two sites and hence two equations are needed. Designating the symbols $\bullet$ and $\circ$ to represent occupied ($\eta_i=1$) and ($\eta_i=0$) empty sites, the system density $\rho$ corresponds to the one-site probability $\rho=P(\scalebox{0.85}{$\bullet$})$. Another quantity to be considered here is the two-site correlation given by $u=P(\scalebox{0.85}{$\circ$}\scalebox{0.85}{$\bullet$})$. From the above model rules, it follows that \begin{equation} \frac{d\rho}{dt}=2P(\scalebox{0.85}{$\circ$}\scalebox{0.85}{$\bullet$}\scalebox{0.85}{$\bullet$} \scalebox{0.85}{$\circ$} \scalebox{0.85}{$\circ$})+P(\scalebox{0.85}{$\circ$}\scalebox{0.85}{$\bullet$}\scalebox{0.85}{$\circ$} \scalebox{0.85}{$\bullet$} \scalebox{0.85}{$\circ$})+3P(\scalebox{0.85}{$\circ$}\scalebox{0.85}{$\bullet$}\scalebox{0.85}{$\circ$} \scalebox{0.85}{$\bullet$} \scalebox{0.85}{$\bullet$})+P(\scalebox{0.85}{$\circ$}\scalebox{0.85}{$\bullet$}\scalebox{0.85}{$\bullet$} \scalebox{0.85}{$\bullet$} \scalebox{0.85}{$\bullet$})-\alpha P(\scalebox{0.85}{$\bullet$}), \end{equation} and \begin{eqnarray} \label{D1} \frac{du}{dt}&=&(1-D)[-\frac{3}{2}P(\scalebox{0.85}{$\circ$}\scalebox{0.85}{$\bullet$}\scalebox{0.85}{$\circ$} \scalebox{0.85}{$\bullet$} \scalebox{0.85}{$\bullet$})-P(\scalebox{0.85}{$\circ$}\scalebox{0.85}{$\bullet$}\scalebox{0.85}{$\bullet$} \scalebox{0.85}{$\bullet$}\scalebox{0.85}{$\bullet$})\\[0.2cm]\nonumber &&-\alpha\,P(\scalebox{0.85}{$\circ$}\scalebox{0.85}{$\bullet$}) +\alpha\,P(\scalebox{0.85}{$\bullet$}\scalebox{0.85}{$\bullet$})] +D[6P(\scalebox{0.85}{$\bullet$}\scalebox{0.85}{$\circ$} \scalebox{0.85}{$\bullet$})-6P(\scalebox{0.85}{$\circ$}\scalebox{0.85}{$\bullet$} \scalebox{0.85}{$\bullet$})],\nonumber \end{eqnarray} for the model A and \begin{equation} \frac{d\rho}{dt}=4P(\scalebox{0.85}{$\circ$}\scalebox{0.85}{$\bullet$}\scalebox{0.85}{$\bullet$} \scalebox{0.85}{$\circ$} \scalebox{0.85}{$\circ$})+2P(\scalebox{0.85}{$\circ$}\scalebox{0.85}{$\bullet$}\scalebox{0.85}{$\circ$} \scalebox{0.85}{$\bullet$} \scalebox{0.85}{$\circ$})+4P(\scalebox{0.85}{$\circ$}\scalebox{0.85}{$\bullet$}\scalebox{0.85}{$\circ$} \scalebox{0.85}{$\bullet$} \scalebox{0.85}{$\bullet$})+P(\scalebox{0.85}{$\circ$}\scalebox{0.85}{$\bullet$}\scalebox{0.85}{$\bullet$} \scalebox{0.85}{$\bullet$} \scalebox{0.85}{$\bullet$})-\alpha P(\scalebox{0.85}{$\bullet$}), \end{equation} \begin{eqnarray} \label{D2} \frac{du}{dt}&=&(1-D)[-2P(\scalebox{0.85}{$\circ$}\scalebox{0.85}{$\bullet$}\scalebox{0.85}{$\circ$} \scalebox{0.85}{$\bullet$} \scalebox{0.85}{$\bullet$})-P(\scalebox{0.85}{$\circ$}\scalebox{0.85}{$\bullet$}\scalebox{0.85}{$\bullet$} \scalebox{0.85}{$\bullet$}\scalebox{0.85}{$\bullet$})\\[0.2cm]\nonumber &&-\alpha\,P(\scalebox{0.85}{$\circ$}\scalebox{0.85}{$\bullet$})+\alpha\,P(\scalebox{0.85}{$\bullet$}\scalebox{0.85}{$\bullet$})] +D[6P(\scalebox{0.85}{$\bullet$}\scalebox{0.85}{$\circ$} \scalebox{0.85}{$\bullet$})-6P(\scalebox{0.85}{$\circ$}\scalebox{0.85}{$\bullet$} \scalebox{0.85}{$\bullet$})],\nonumber \end{eqnarray} for the model B. Here the symbol $P(\eta_0,\eta_1,\eta_2,\eta_3,\eta_4)$ denotes the probability of finding the central site in the state $\eta_0$ and its four nearest neighbors in the states $\eta_1$, $\eta_2$, $\eta_3$ and $\eta_4$. The pair mean-field approximation consists of rewriting the $n$-site probabilities ($n>2$) as products of two-site in such a way that \begin{equation}\label{pair_approx} P(\eta_0,\eta_1,...,\eta_{n-1}) \simeq \frac{P(\eta_0,\eta_1) P(\eta_0,\eta_2)...P(\eta_0,\eta_{n-1})}{P(\eta_0)^{n-2}}\,. \end{equation} From this approximation, the above equations read \begin{equation} \frac{d\rho}{dt}=\frac{3u^{2}}{(1-\rho)}-\frac{3u^{3}}{(1-\rho)^{2}}+ \frac{u^{4}}{(1-\rho)^{3}}-\alpha\rho, \end{equation} \begin{eqnarray} \frac{du}{dt}&=&(1-D)\left[-\frac{3u^{3}}{2(1-\rho)^{2}}+\frac{u^{4}}{2(1-\rho)^{3}} -2\alpha\,u+\alpha\rho\right]+\nonumber\\ [0.2cm] &&+D\left[6u-\frac{6u^{2}}{\rho(1-\rho)}\right], \end{eqnarray} for the model A and \begin{equation} \frac{d\rho}{dt}=\frac{6u^{2}}{(1-\rho)}-\frac{8u^{3}}{(1-\rho)^{2}}+ \frac{3u^{4}}{(1-\rho)^{3}}-\alpha\rho, \end{equation} \begin{eqnarray} \frac{du}{dt}&=&(1-D)\left[-\frac{2u^{3}}{(1-\rho)^{2}}+\frac{u^{4}}{(1-\rho)^{3}} -2\alpha\,u+\alpha\rho\right]+\nonumber\\ [0.2cm] &&+D\left[6u-\frac{6u^{2}}{\rho(1-\rho)}\right], \end{eqnarray} for the model B. The steady solutions are obtained by taking $\frac{d\rho}{dt}=\frac{du}{dt}=0$ implying that for locating the transition point and the order of transition it is required to solve a system of two coupled equations for a given set of parameters ($\alpha$,\,$D$). Although alternative treatments have been considered \cite{evans3}, here we shall identify the order of transitions by inspecting the dependence of $\rho$ vs $\alpha$. In similarity with equilibrium transitions, the existence of a spinodal behavior (with $\rho$ increasing by raising $\alpha$) signals a discontinuous transition. Despite this, no analogous treatments similar to the ``Maxwell construction'' are available. For instance, the coexistence points have been estimated by the maximum value of $\alpha$ and the spinodal behavior replaced by a jump in $\rho$. Fig. 1 $(a)$ and $(b)$ show the phase diagram for distinct diffusion rates. For all values of $D$, mean-field results predict discontinuous transitions separating the absorbing and the active phases. However, as $D$ increases, the transition point moves to larger values of $\alpha_0$ and the active phase becomes less dense. For example, for $D=0.1$ and $D=0.9$, the active phases have densities $\rho_{ac}=0.445$ and $0.371$ (model A) and $\rho_{ac}=0.419$ and $0.324$ (model B), respectively. For $D \rightarrow 1$, one recovers the limit of fully uncorrelated particles, so that $P(\eta_0,\eta_1,...,\eta_{n-1})=P(\eta_0) P(\eta_1) ...P(\eta_{n-1})$. In this limit, Eqs. (1) and (3) become equivalent to those obtained from the one-site MFT given by $\frac{d \rho}{dt}=\rho^2(1-\rho)[3-3\rho+\rho^2]-\alpha \rho$ and $\frac{d \rho}{dt}=\rho^2(1-\rho)[6-8\rho+3\rho^2]-\alpha \rho$, for the models A and B, respectively. For such regime, transition points take place at $\alpha_0=0.4724$ with active phase density $\rho_{ac}=0.370$ (model A) and $\alpha_0=0.8154$ with $\rho_{ac}=0.322$ (model B). Thus, despite the increase of diffusion displaces particles, MFT predicts a discontinuous transition, regardless the strength of the diffusion rate. As a difference between models, active phases are somewhat more dense for model A than in model B. \begin{figure}[h] \centering \includegraphics[scale=0.37]{diagrams_dif_MFT.eps} \caption{Two-site mean field phase diagrams for diffusive model versions A $(a)$ and B $(b)$. Dashed lines denote discontinuous phase transitions. The black circles indicate the $D \rightarrow 1$ limit predicted by the one site MFT. Insets show the two-site MFT results for $D=0.1$ and $D=0.9$. Dotted lines correspond to spinodal behaviors, being replaced here by a jump.} \label{fig1} \end{figure} \section{Numerical results} Numerical simulations have been performed for distinct system sizes of a square lattice with periodic boundary conditions. Due to the lack of a general theory for discontinuous absorbing transitions, including the absence of a finite-size scaling (FSS) theory and unknown spreading experiments behavior for $d \ge 2$, two sort of analysis will be presented. In the former, we study the time decay of the density $\rho$ starting from a fully occupied initial condition for distinct independent runs. As for critical and discontinuous phase transitions, for small $\alpha$ the density $\rho$ converges to a definite value indicating endless activity, in which particles are continuously created and annihilated. On the contrary, for sufficiently large $\alpha$'s, the system density $\rho$ vanishes exponentially toward a complete particle extinction. The ``coexistence'' point $\alpha_{0}$ is the separatrix between above regimes, whereas at the critical point $\alpha_c$ the density $\rho$ vanishes algebraically following a power-law behavior $\rho \sim t^{-\theta}$, with $\theta$ its associated critical exponent. For the DP universality class it reads $\theta=0.4505(10)$ \cite{henkel}. Thus, the difference of above behaviors will be used to identify the order of phase transition. Second, the behavior of typical quantities in the steady regime is investigated. For instance, we apply the models dynamics together with the quasi-steady method \cite{qs}. Briefly, it consists of storing a list of $M$ active configurations (here we store $M=2000-3000$ configurations) and whenever the system falls into the absorbing state a configuration is randomly extracted from the list. The ensemble of stored configurations is continuously updated, where for each MC step a configuration belonging to the list is replaced with probability ${\tilde p}$ (typically one takes ${\tilde p}=0.01$) by the actual system configuration, provided it is not absorbing. Discontinuous transitions are typically signed by bimodal probability distribution $P_\rho$ (characterizing the ``coexistence'' between absorbing and active phases) and for finite systems, a peak in the order-parameter variance $\chi=L^{2}[\langle \rho^{2}\rangle-\langle \rho \rangle^{2}]$ is expected close to the transition point. For equilibrium systems, the maximum of $\chi$ and other quantities scale with the system volume and its position $\alpha_L$ obeys the asymptotic relation $\alpha_L =\alpha_0 - c/L^2$ \cite{rBoKo}, being $\alpha_0$ the transition point in the thermodynamic limit and $c$ a constant. Although the finite size scaling properties are unknown for nonequilibrium systems, results for some first-order transitions into an absorbing state have shown a similar scaling than the equilibrium case (with the system volume) \cite{carlos,sina}. Results for the model A and three representative diffusion rates ($D=0.1$, $D=0.5$ (not shown) and $0.9$) are summarized in Figs. \ref{fig2} and \ref{fig4}. In all cases, panels $(a)$ show that there is a threshold point $\alpha_0$ separating an active state (signed by the convergence to a definite value of $\rho$) from an exponential vanishing of $\rho$. Such values increase by raising $D$, yielding at $\alpha_0 \sim 0.2590$, $0.360$ and $0.436$ for $D=0.1$, $0.5$ and $0.9$, respectively. Since no power-law behavior is presented, such analysis provides the first evidence of a discontinuous transition for all diffusion strengths. In order to confirm this, panels $(b)$ show the probability distribution $P_{\rho}$ for distinct system sizes. In all cases $P_{\rho}$ presents a bimodal shape, whose position of equal peaks deviate mildly as $L$ increases. Also, the peaks corresponding to the active and absorbing phases present distinct dependencies on $L$. Whereas active phase densities $\rho_{ac}$ converge to well defined values, $\rho_{ab}$ vanishes with $1/L^2$ (insets). For $D=0.1,0.5$ and $0.9$ the $\rho_{ac}$'s converge to $0.603(1)$, $0.501(2)$ and $0.434(2)$, respectively. These features are similar to the restrictive models studied in Ref. \cite{carlos}. For $D=0.99$ (not shown) a bimodal probability distribution with active phase centered at $\rho_{ac}=0.405(1)$ is observed. In the third analysis, the behavior of the system density $\rho$ and its variance $\chi$ is presented. Note that $\rho$ vanishes in a short range of $\alpha$ followed by a peak of the variance $\chi$ (panels $(c)$ and their insets). For each $D$, the positions $\alpha_L$'s, in which $\chi$ presents a maximum, scale with $L^{-2}$ whereas the maximum of $\chi$ increases with $L^2$ (panels $(d)$ and insets). These features are similar to equilibrium discontinuous phase transitions \cite{rBoKo,fioreprl} and with the scalings in Ref. \cite{carlos}. From this behavior, we obtain the values $\alpha_0=0.2600(1)$ ($D=0.1$), $0.3624(2)$ ($D=0.5$) and $0.442(1)$ ($D=0.9$), which agree with previous estimates, obtained from the time decay of $\rho$. Small discrepancies between estimates can be attributed to the lattice simulated for the time decay be finite, due to the uncertainties in the position of peaks or both. Thus, steady analysis reinforces above conclusions concerning the phase transition be discontinuous regardless the diffusion rate. \begin{figure}[h] \centering \includegraphics[scale=0.35]{fig2_b.eps} \caption{For the model A and $D=0.1$, panel $(a)$ shows the time evolution of $\rho$ for distinct $\alpha$'s and $L=150$. In $(b)$ the quasi-stationary probability distribution $P_{\rho}$ for $L$'s in which the peaks have the same height. In the inset, the log-log plot of the quasi-steady densities vs L. In $(c)$ $\chi$ and $\rho$ (inset) vs $\alpha$ for distinct system sizes $L$. In $(d)$, the scaling plot of $\alpha_{L}$, in which $\chi$ is maximum, vs $L^{-2}$. Inset shows the log-log plot of the maximum of $\chi$ vs $L$ and the straight line has slope $2$.} \label{fig2} \end{figure} \begin{figure}[h] \centering \includegraphics[scale=0.35]{fig4_b.eps} \caption{For the model A and $D=0.9$, panel $(a)$ shows the time evolution of $\rho$ for distinct $\alpha$'s and $L=150$. In $(b)$ the quasi-stationary probability distribution $P_{\rho}$ for $L$'s in which the peaks have the same height. In the inset, the log-log plot of the quasi-steady densities vs L. In $(c)$ $\chi$ and $\rho$ (inset) vs $\alpha$ for distinct system sizes $L$. In $(d)$, the scaling plot of $\alpha_{L}$, in which $\chi$ is maximum, vs $L^{-2}$. Inset shows the log-log plot of the maximum of $\chi$ vs $L$ and the straight line has slope $2$.} \label{fig4} \end{figure} Next, we extended above analysis for model B, with results summarized in Figs. \ref{fig42} and \ref{fig43} for $D=0.1$ and $D=0.9$, respectively. As for the model A, both values of $D$ show that the separatrix between active and absorbing regimes are signed by the absence of a power-law behavior and thus the transitions seem to be discontinuous. In particular, the separatrix points (panels $(a)$) yield close to $\alpha_0 \sim 0.5015$ ($D=0.1$) and $0.770$ ($D=0.9$). In order to confirm, we also examine the probability distribution as well as the behavior of $\rho$ and $\chi$. In the former, $P_{\rho}$ also presents bimodal shapes exhibiting two well defined peaks (panels $(b)$). Whenever $\rho_{ac}$'s (insets) saturate as $L$ increases, the $\rho_{ab}$'s vanish as $1/L^2$. In addition, active phase also becomes less dense as $D$ increases, reading $\rho_{ac}=0.482(1)$ and $0.372(1)$ for $D=0.1$ and $0.9$, respectively. As for the model A and those studied in Ref. \cite{carlos} the positions of peaks $\alpha_L$'s (in which $\chi$ presents a peak) as well the peaks also scale with $L^{-2}$ and $L^{2}$ (panels $(d)$), respectively. From this scaling behavior, we obtain the values $\alpha_0=0.5027(2)$ and $0.7844(2)$ for $D=0.1$ and $D=0.9$, respectively, in agreement with above estimates (from the time decay of $\rho$). \begin{figure}[h] \centering \includegraphics[scale=0.35]{fig1_modelb.eps} \caption{For the model B and $D=0.1$, panel $(a)$ shows the time evolution of $\rho$ for distinct $\alpha$'s and $L=150$. In $(b)$ the quasi-stationary probability distribution $P_{\rho}$ for $L$'s in which the peaks have the same height. In the inset, the log-log plot of the quasi-steady densities vs L. In $(c)$ $\chi$ and $\rho$ (inset) vs $\alpha$ for distinct system sizes $L$. In $(d)$, the scaling plot of $\alpha_{L}$, in which $\chi$ is maximum, vs $L^{-2}$. Inset shows the log-log plot of the maximum of $\chi$ vs $L$ and the straight line has slope $2$.} \label{fig42} \end{figure} \begin{figure}[h] \centering \includegraphics[scale=0.345]{fig2_modelb.eps} \caption{For the model B and $D=0.9$, panel $(a)$ shows the time evolution of $\rho$ for distinct $\alpha$'s and $L=150$. In $(b)$ the quasi-stationary probability distribution $P_{\rho}$ for $L$'s in which the peaks have the same height. In the inset, the log-log plot of the quasi-steady densities vs L. In $(c)$ $\chi$ and $\rho$ (inset) vs $\alpha$ for distinct system sizes $L$. In $(d)$, the scaling plot of $\alpha_{L}$, in which $\chi$ is maximum, vs $L^{-2}$. Inset shows the log-log plot of the maximum of $\chi$ vs $L$ and the straight line has slope $2$.} \label{fig43} \end{figure} Thus, in similarity with MFT, numerical simulations suggest that the diffusion does not change the order of phase transition, although clusters become less compact as $D$ increases. Similar conclusions are found for extremely low diffusion strengths. For example, for the model A with $D=0.01$ and $0.05$, all above features are verified at $\alpha \sim 0.215$ and $0.239$, respectively. Moreover, results of both models contrast partially with those obtained by Villa-Martin et al. \cite{munoz214} in the regime of low diffusion rates, although they agree qualitatively in the limit of intermediate and large diffusion regimes. A possible explanation for such differences is presented as follows: As shown in Ref. \cite{munoz214}, a coarse-grained description of a discontinuous absorbing transition is the differential equation $\partial_{t}\rho=-\alpha\rho-b\rho^2-c\rho^3 +D\nabla^{2}\rho+ \eta(x,t)$, where $D\nabla^2\rho$ corresponds to the diffusion term and $\eta(x,t)$ is the (white) noise. The parameter $b$ is responsible for the density discontinuity, since their signs $b<0$ $(>0)$ provide two (one) stable solutions. On the other hand, the parameter $c$ is required to be positive ($c>0$) for ensuring finite densities. The deterministic part of above equation ($\alpha \rho- b \rho^2- c\rho^3$) can be obtained for example by taking a fully connected lattice with the second Sch\"ogl transition rates $W^{-}(\rho \rightarrow \rho-1/L^2)=\alpha\rho$ and $W^{+}(\rho \rightarrow \rho+1/L^2)=\rho^2(1-\rho)$, where at each time instant the system density $\rho$ changes by a factor $\pm 1/L^2$. A similar reasoning can be extended for the present studied models, but with different $W^+$'s. In particular, they read $W^{+}(\rho \rightarrow \rho+1/L^2)=\rho^2(1-\rho)(3-3\rho+\rho^2)$ and $W^{+}(\rho \rightarrow \rho+1/L^2)=\rho^2(1-\rho)(6-8\rho+3\rho^2)$ for the models A and B, respectively and leading to the (deterministic) terms $\partial_{t}\rho=-\alpha\rho+3\rho^2-6\rho^3+4\rho^4-\rho^5$ (model A) and $\partial_{t}\rho=-\alpha\rho+6\rho^2-14\rho^3+11\rho^4-3\rho^5$ (model B). Thus, different coarse grained descriptions can explain the difference between results in the regime of low diffusion rates. It is worth remarking that further studies are still required to confirm above points. Despite the similarities between models A and B, some differences are clearly observed. In particular, as predicted by the MFT, compact clusters are somewhat less dense for model B than for model A. Also, the dependence between $D$ and $\alpha$ in both models is somewhat different. Extending the aforementioned analysis for distinct values of $D$, we obtain the phase diagram shown in Fig. \ref{fig5}. For both models, in the limit $D \rightarrow 1$ the transition points approach their values predicted by the MFT. \begin{figure}[h] \centering \includegraphics[scale=0.4]{diagrams_dif_MC.eps} \caption{The phase diagram in the plane $D-\alpha$ obtained from MC simulations for the models A (left) and B (right). Dashed line denotes discontinuous phase transitions.} \label{fig5} \end{figure} \section{Conclusions} To sum up, we investigated the influence of diffusion in two simple models presenting discontinuous absorbing phase transitions. They have been studied via mean-field analysis and distinct numerical simulations. All results suggest that transitions are discontinuous irrespective the strength of the diffusion, in contrast to recent findings in which limited diffusion induces a critical transition \cite{munoz214}. Thus our results indicate not only an additional feature of diffusion, but also the possible differences between lattice model and Langevin approaches. It is worth emphasizing that the study of other models variants (taking into account the inclusion of distinct annihilation rates) are required for checking whether the disagreement between approaches is also verified in other cases. Other remarkable result is that all transitions presented a finite-size scaling (with the system volume) similar to the discontinuous transitions studied in Ref. \cite{carlos}. Although further investigations are still required, the obtained results reinforces the possibility of a general scaling for nonequilibrium phase transitions. \section*{ACKNOWLEDGMENT} The authors wish to thank Brazilian scientific agencies CNPq, INCT-FCx for the financial support. \section*{References}
\section{Introduction}\label{sec:introduction} In this paper we propose a randomized algorithm for quick detection of high-degree entities in large online social networks. The entities can be, for example, users, interest groups, user categories, geographical locations, etc. For instance, one can be interested in finding a list of Twitter users with many followers or Facebook interest groups with many members. The importance of this problem is attested by a large number of companies that continuously collect and update statistics about popular entities in online social networks ({\it twittercounter.com}, {\it followerwonk.com}, {\it twitaholic.com}, {\it www.insidefacebook.com}, {\it yavkontakte.ru} just to name a few). The problem under consideration may seem trivial if one assumes that the network structure and the relation between entities are known. However, even then finding for example the top-$k$ in-degree nodes in a directed graph $G$ of size $N$ takes the time ${\rm O}(N)$. For very large networks, even linear complexity is too high cost to pay. Furthermore, the data of current social networks is typically available only to managers of social networks and can be obtained by other interested parties only through API (Application Programming Interface) requests. API is a set of request messages, along with a definition of the structure of response messages. Using one API request it is usually possible to discover either friends of one given user, or his/her interest groups, or the date when his/her account was created, etc. The rate of allowed API requests is usually very limited. For instance, Twitter has the limit of one access per minute for one standard API account (see \textit{dev.twitter.com}). Then, in order to crawl the entire network with a billion users, using one standard API account, one needs more than 1900 years. Hence currently, there is a rapidly growing interest in algorithms that evaluate specific network properties, using only local information (e.g., the degree of a node and its neighbors), and give a good approximate answer in the number of steps that is sublinear in the network size. Recently, such algorithms have been proposed for PageRank evaluation~\cite{Avrachenkov2011Top-kPPR,SublinearPageRank,Borgs2013PageRank}, for finding high-degree nodes in graphs~\cite{Avrachenkov2012Top-k,Brautbar2010high_degree,Cooper2012high_degree,Kumar2008}, and for finding the root of a preferential attachment tree~\cite{Borgs2012pa_root}. In this paper, we propose a new two-stage method for finding high-degree nodes in large directed networks with highly skewed in-degree distribution. We demonstrate that our algorithm outperforms other known methods by a large margin and has a better precision than the for-profit Twitter statistics {\it twittercounter.com}. \section{Problem formulation and our contribution} Let $V$ be a set of $N$ entities, typically users, that can be accessed using API requests. Let $W$ be another set of $M$ entities (possibly equal to $V$). We consider a bipartite graph $(V,W,E)$, where a directed edge $(v,w)\in E$, with $v\in V$, and $w\in W$, represents a relation between $v$ and $w$. In our particular model of the Twitter graph $V$ is a set of Twitter users, $W=V$, and $(v,w)\in E$ means that $v$ follows $w$ or that $v$ retweeted a tweet of $w$. Note that any directed graph $G=(V,E)$ can be represented equivalently by the bipartite graph $(V,V,E)$. One can also suppose that $V$ is a set of users and $W$ is a set of interest groups, while the edge $(v,w)$ represents that the user $v$ belongs to the group $w$. Our goal is to quickly find the top in-degree entities in $W$. In this setting, throughout the paper, we use the terms `nodes', `vertices', and `entities' interchangeably. We propose a very simple and easy-to-implement algorithm that detects popular entities with high precision using a surprisingly small number of API requests. Most of our experiments are performed on the Twitter graph, because it is a good example of a huge network (approximately a billion of registered users) with a very limited rate of requests to API. We use only 1000 API requests to find the top-100 Twitter users with a very high precision. We also demonstrate the efficacy of our approach on the popular Russian online social network VKontakte (\textit{vk.com}) with more than 200 million registered users. We use our algorithm to quickly detect the most popular interest groups in this social network. Our experimental analysis shows that despite of its simplicity, our algorithm significantly outperforms existing approaches, e.g., \cite{Avrachenkov2012Top-k,Brautbar2010high_degree,Kumar2008}. Moreover, our algorithm can be used in a very general setting for finding the most popular entities, while some baseline algorithms can only be used for finding nodes of largest degrees in directed \cite{Kumar2008} or undirected \cite{Avrachenkov2012Top-k} graphs. In most social networks the degrees of entities show great variability. This is often modeled using power laws, although it has been often argued that the classical Pareto distribution does not always fit the observed data. In our analysis we assume that the incoming degrees of the entities in $W$ are independent random variables following a {\it regularly varying} distribution $G$: \begin{equation}\label{eq:regular} 1-G(x)=L(x)x^{-1/\gamma},\quad x>0,\; \gamma>0, \end{equation} where $L(\cdot)$ is a slowly varying function, that is, \[\lim_{x\to\infty}L(tx)/L(x)=1,\quad t>0.\] $L(\cdot)$ can be, for example, a constant or logarithmic function. We note that \eqref{eq:regular} describes a broad class of heavy-tailed distributions without imposing the rigid Pareto assumption. An important contribution of this work is a novel analysis of the proposed algorithm that uses powerful results of the Extreme Value Theory (EVT)~--- a branch of probability that studies extreme events and properties of high order statistics in random samples. We refer to \cite{deHaan-Ferreira} for a comprehensive introduction to EVT. Using EVT we can accurately predict the average fraction of correctly identified top-$k$ nodes and obtain the algorithm's complexity in terms of the number of nodes in $V$. We show that the complexity is sublinear if the in-degree distribution of the entities in $W$ is heavy tailed, which is usually the case in real networks. The rest of the paper is organized as follows. In Section~\ref{sec:literature}, we give a short overview of related work. We formally describe our algorithm in Section~\ref{sec:algorithm}, then we introduce two performance measures in Section~\ref{sec:criteria}. Section~\ref{sec:experiments} contains extensive experimental results that demonstrate the efficiency of our algorithm and compare it to baseline strategies. In Sections~\ref{sec:analysis}-\ref{sec:complexity} we present a detailed analysis of the algorithm and evaluate its optimal parameters with respect to the two performance measures. Section~\ref{sec:conclusion} concludes the paper. \section{Related work}\label{sec:literature} Over the last years data sets have become increasingly massive. For algorithms on such large data any complexity higher than linear (in dataset size) is unacceptable and even linear complexity may be too high. It is also well understood that an algorithm which runs in sublinear time cannot return an exact answer. In fact, such algorithms often use randomization, and then errors occur with positive probability. Nevertheless, in practice, a rough but quick answer is often more valuable than the exact but computationally demanding solution. Therefore, sublinear time algorithms become increasingly important and many studies of such algorithms appeared in recent years (see, e.g., \cite{Sublinear1,Sublinear2,Sublinear3,Sublinear4}). An essential assumption of this work is that the network structure is not available and has to be discovered using API requests. This setting is similar to on-line computations, where information is obtained and immediately processed while crawling the network graph (for instance the World Wide Web). There is a large body of literature where such on-line algorithms are developed and analyzed. Many of these algorithms are developed for computing and updating the PageRank vector~\cite{Abiteboul,MonteCarloAvrachenkov,SublinearPageRank,Fagaras}. In particular, the algorithm recently proposed in \cite{SublinearPageRank} computes the PageRank vector in sublinear time. Furthermore, probabilistic Monte Carlo methods \cite{MonteCarloAvrachenkov,Bahmani2010PRMonteCarlo,Fagaras} allow to continuously update the PageRank as the structure of the Web changes. Randomized algorithms are also used for discovering the structure of social networks. In \cite{Leskovec2006sampling} random walk methods are proposed to obtain a graph sample with similar properties as a whole graph. In \cite{Gjoka} an unbiased random walk, where each node is visited with equal probability, is constructed in order to find the degree distribution on Facebook. Random walk based methods are also used to analyse Peer-to-Peer networks \cite{Massoulie2006}. In \cite{Borgs2012pa_root} traceroute algorithms are proposed to find the root node and to approximate several other characteristics in a preferential attachment graph. The problem of finding the most popular entities in large networks based only on the knowledge of a neighborhood of a current node has been analyzed in several papers. A random walk algorithm is suggested in \cite{Cooper2012high_degree} to quickly find the nodes with high degrees in a preferential attachment graph. In this case, transitions along undirected edges $x,y$ are proportional to $(d(x)d(y))^b$, where $d(x)$ is the degree of a vertex $x$ and $b > 0$ is some parameter. In \cite{Avrachenkov2012Top-k} a random walk with restart that uses only the information on the degree of a currently visited node was suggested for finding large degree nodes in undirected graphs. In \cite{Brautbar2010high_degree} a local algorithm for general networks, power law networks, and preferential attachment graphs is proposed for finding a node with degree, which is smaller than the maximal by a factor at most $c$. Another crawling algorithm \cite{Kumar2008} is proposed to efficiently discover the correct set of web pages with largest incoming degrees in a fixed network and to track these pages over time when the network is changing. Note that the setting in \cite{Kumar2008} is different from ours in several aspects. For example, in our case we can use API to inquire the in-degree of any given item, while in the World Wide Web the information on in-links is not available, the crawler can only observe the in-links that come from the pages already crawled. In Section~\ref{sec:baselines} we show that our algorithm outperforms the existing methods by a large margin. Besides, several of the existing methods such as the ones in \cite{Avrachenkov2012Top-k} and \cite{Kumar2008} are designed specifically to discover the high degree nodes, and they cannot be easily adapted for other tasks, such as finding the most popular user categories or interest groups, while the algorithm proposed in this paper is simpler, much faster, and more generic. To the best of our knowledge, this is the first work that presents and analyzes an efficient algorithm for retrieving the most popular entities under realistic API constraints. \section{Algorithm description}\label{sec:algorithm} Recall that we consider a bipartite graph $(V,W,E)$, where $V$ and $W$ are sets of entities and $(v,w)\in E$ represents a relation between the entities. Let $n$ be the allowed number of requests to API. Our algorithm consists of two steps. We spend $n_1$ API requests on the first step and $n_2$ API requests on the second step, with $n_1+n_2=n$. See Algorithm~\ref{algo1} for the pseudocode. \IncMargin{1em} \begin{algorithm} \SetKwInOut{Input}{input}\SetKwInOut{Output}{output} \Input{Set of entities $V$ of size $N$, set of entities $W$ of size $M$, number of random nodes $n_1$ to select from $V$, number of candidate nodes $n_2$ from $W$} \Output{Nodes $w_1, \dots w_{n_2}\in W$, their degrees $d_1, \dots, d_{n_2}$} \BlankLine \For{$w$ in $W$}{ $S[w] \leftarrow 0$\;} \For{$i\leftarrow 1$ \KwTo $n_1$}{ $v \leftarrow random(N)$\; \ForEach{$w$ in $OutNeighbors(v) \subset W$}{ $S[w] \leftarrow S[w]+1$\; }} $w_1, \dots, w_{n_2} \leftarrow Top\_n_2(S)$ // $S[w_1], \ldots, S[w_{n_2}]$ are the top $n_2$ maximum values in $S$\; \For{$i\leftarrow 1$ \KwTo $n_2$}{ $d_i \leftarrow InDegree(w_i)$\;} {\caption{Two-stage algorithm} \label{algo1}} \end{algorithm}\DecMargin{1em} \textbf{First stage.} We start by sampling uniformly at random a set $A$ of $n_1$ nodes $v_1, \ldots, v_{n_1}\in V$. The nodes are sampled independently, so the same node may appear in $A$ more than once, in which case we regard each copy of this node as a different node. Note that multiplicities occur with a very small probability, approximately $1-e^{-{n_1^2}/{(2N)}}$. For each node in $A$ we record its out-neighbors in $W$. In practice, we bound the number of recorded out-links by the maximal number of IDs that can be retrieved within one API request, thus the first stage uses exactly $n_1$ API requests. For each $w\in W$ we identify $S[w]$, which is the number of nodes in $A$ that have a (recorded) edge to $w$. \textbf{Second stage.} We use $n_2$ API requests to retrieve the actual in-degrees of the $n_2$ nodes with the highest values of $S[w]$. The idea is that the nodes with the largest in-degrees in $W$ are likely to be among the $n_2$ nodes with the largest $S[w]$. For example, if we are interested in the top-$k$ in-degree nodes in a directed graph, we hope to identify these nodes with high precision if $k$ is significantly smaller than $n_2$. \section{Performance metrics} \label{sec:criteria} The main constraint of Algorithm~\ref{algo1} is the number of API requests we can use. Below we propose two performance metrics: the average fraction of correctly identified top-$k$ nodes and the first-error index. We number the nodes in $W$ in the deceasing order of their in-degrees and denote the corresponding in-degrees by $F_1\ge F_2\ge \cdots\ge F_M$. We refer to $F_j$ as the $j$-th order statistic of the in-degrees in $W$. Further, let $S_j$ be the number of neighbors of a node $j$, $1 \le j \le M$, among the $n_1$ randomly chosen nodes in $V$, as described in Algorithm~\ref{algo1}. Finally, let $S_{i_1}\ge S_{i_2}\ge\ldots\ge S_{i_M}$ be the order statistics of $S_1,\ldots,S_M$. For example, $i_1$ is the node with the largest number of neighbors among $n_1$ randomly chosen nodes, although $i_1$ may not have the largest degree. Clearly, node $j$ is identified if it is in the set $\{i_1, i_2,\ldots, i_{n_2}\}$. We denote the corresponding probability by \begin{equation}\label{eq:pj} P_j(n_1):=\mathrm{P}(j\in \{i_1,\ldots,i_{n_2}\})\,. \end{equation} The first performance measure is the average fraction of correctly identified top-$k$ nodes. This is defined in the same way as in~\cite{Avrachenkov2011Top-kPPR}: \begin{align} \nonumber \mathbb{E}[\mbox{fraction of correctly identified top-$k$ entities}] &\\ = \frac{1}{k}\sum_{j=1}^k P_j(n_1).& \label{eq:prediction} \end{align} The second performance measure is the first-error index, which is equal to $i$ if the top $(i-1)$ entities are identified correctly, but the top-$i$th entity is not identified. If all top-$n_2$ entities are identified correctly, we set the first-error index equal to $n_2+1$. Using the fact that for a discrete random variable $X$ with values $1,2,\ldots,K+1$ holds $\mathbb{E}(X)=\sum_{j=1}^{K+1}\mathrm{P}(X\ge j)$, we obtain the average first-error index as follows: \begin{align} \nonumber \mathbb{E}[\mbox{1st-error index}]&=\sum_{j=1}^{n_2+1}\mathrm{P}(\mbox{1st-error index}\ge j)\\ &= \sum_{j=1}^{n_2+1}\prod_{l=1}^{j-1}P_l(n_1). \label{eq:prediction1} \end{align} If the number $n$ of API requests is fixed, then the metrics \eqref{eq:prediction} and \eqref{eq:prediction1} involve an interesting trade-off between $n_1$ and $n_2$. On the one hand, $n_1$ should be large enough so that the values $S_i$'s are sufficiently informative for filtering out important nodes. On the other hand, when $n_2$ is too small we expect a poor performance because the algorithm returns a top-$k$ list based mainly on the highest values of $S_i$'s, which have rather high random fluctuations. For example, on Figure~\ref{fig:fraction_correct}, when $n_2=k=100$, the algorithm returns the nodes $\{i_1,\ldots,i_{100}\}$, of which only 75\% belong to the true top-100. Hence we need to find the balance between $n_1$ and $n_2$. This is especially important when $n$ is not very large compared to $k$ (see Figure~\ref{fig:fraction_correct} with $n=1000$ and $k=250$). \section{Experiments}\label{sec:experiments} This section is organized as follows. First, we analyze the performance of our algorithm (most of the experiments are performed on the Twitter graph, but we also present some results on the CNR-2000 graph). Then we compare our algorithm with baseline strategies on the Twitter graph and show that the algorithm proposed in this paper significantly outperforms existing approaches. Finally, we demonstrate another application of our algorithm by identifying the most popular interest groups in the large online social network VKontakte. All our experiments are reproducible: we use public APIs of online social networks and publicly available sample of a web graph. \subsection{Performance of the proposed algorithm} First, we show that our algorithm quickly finds the most popular users in Twitter. Formally, $V$ is a set of Twitter users, $W=V$, and $(v,w)\in E$ iff $v$ is a follower of $w$. Twitter is an example of a huge network with a very limited access to its structure. Information on the Twitter graph can be obtained via Twitter public API. The standard rate of requests to API is one per minute (see \textit{dev.twitter.com}). Every vertex has an ID, which is an integer number starting from $12$. The largest ID of a user is $\sim 1500$M (at the time when we performed the experiments). Due to such ID assignment, a random user in Twitter can be easily chosen. Some users in this range have been deleted, some are suspended, and therefore errors occur when addressing the IDs of these pages. In our implementation we skip errors and assume that we do not spend resources on such nodes. The fraction of errors is approximately $30\%$. In some online social networks the ID space can be very sparse and this makes problematic the execution of uniform sampling in the first stage of our algorithm. In such situation we suggest to use random walk based methods (e.g., Metropolis-Hastings random walk from \cite{Gjoka} or continuous-time random walk from \cite{Massoulie2006}) that produce approximately uniform sampling after a burn-in period. To remove the effect of correlation, one can use a combination of restart \cite{ART10} and thinning \cite{Avrachenkov2012Top-k,Gjoka}. Given an ID of a user, a request to API can return one of the following: i) the number of followers (in-degree), ii) the number of followees (out-degree), or iii) at most 5000 IDs of followers or followees. If a user has more than 5000 followees, then all their IDs can be retrieved only by using several API requests. Instead, as described above, we record only the first 5000 of the followees and ignore the rest. This does not affect the performance of the algorithm because we record followees of randomly sampled users, and the fraction of Twitter users with more than 5000 followees is very small. \begin{figure} \centerline{\includegraphics[width = 0.5\textwidth]{from_top_fraction}} \caption{The fraction of correctly identified top-$k$ most followed Twitter users as a function of $n_2$, with $n=1000$.} \label{fig:fraction_correct} \end{figure} In order to obtain the ground truth on the Twitter graph, we started with a top-1000 list from the publicly available source {\it twittercounter.com}. Next, we obtained a top-1000 list by running our algorithm with $n_1=n_2=20\,000$. We noticed that 1) our algorithm discovers all top-1000 users from {\it twittercounter.com}, 2) some top users identified by our algorithm are not presented in the top-1000 list on {\it twittercounter.com}. Then, we obtained the ground truth for top-1000 users by running our algorithm with ample number of API requests: $n_1 = n_2 = 500\,000$. First we analyzed the fraction of correctly identified top-$k$ nodes (see Equation~\eqref{eq:prediction}). Figure~\ref{fig:fraction_correct} shows the average fraction of correctly identified top-$k$ users for different $k$ over 100 experiments, as a function of $n_2$, when $n=1000$, which is very small compared to the total number of users. Remarkably we can find the top-50 users with very high precision. Note that, especially for small $k$, the algorithm has a high precision in a large range of parameters. We also looked at the first-error index (see Equation~\eqref{eq:prediction1}), i.e., the position of the first error in the top list. Again, we averaged the results over 100 experiments. Results are shown on Figure~\ref{fig:first_mistake} (red line). Note that with only 1000 API requests we can (on average) correctly identify more than 50 users without any omission. \begin{figure} \centerline{\includegraphics[width = 0.5\textwidth]{first_error.pdf}} \caption{The first-error index as a function of $n_2$, with $n=1000$, on Twitter.} \label{fig:first_mistake} \end{figure} \begin{figure} \centerline{\includegraphics[width = 0.5\textwidth]{from_top_CNR}} \caption{The fraction of correctly identified top-$k$ in-degree nodes in the CNR-2000 graph as a function of $n_2$, with $n=1000$.} \label{fig:CNR} \end{figure} Although in this paper we mostly focus on the Twitter graph (since it is a huge network with a very limited rate of requests to API), we also demonstrated the performance of our algorithm on CNR-2000 graph ({\it law.di.unimi.it/webdata/cnr-2000}). This graph is a sample of the Italian CNR domain. It is much smaller and there are no difficulties in obtaining the ground truth here. We get very similar results for this graph (see Figure~\ref{fig:CNR}). Interestingly, the performance of the algorithm is almost insensitive to the network size: the algorithm performs similarly on the network with a billion nodes as on the network with half a million nodes. \subsection{Comparison with baseline algorithms}\label{sec:baselines} Literature suggests several solutions for the problem studied here. Not every solution is feasible in the setting of a large unknown realistic network. For example, random-walk-based algorithms that require the knowledge of the degrees of all neighbors of a currently visited node, such as the one in \cite{Cooper2012high_degree}, are not applicable. Indeed if we want to make a transition from a vertex of degree $d$, we need at least $d$ requests to decide where to go. So once the random walk hits a vertex of high degree, we may spend all the allowed resources on just one transition of the random walk. In this section, we compare our algorithm with the algorithms suggested in \cite{Avrachenkov2012Top-k}, \cite{Brautbar2010high_degree}, and \cite{Kumar2008}. We start with the description of these algorithms. \textbf{RandomWalk}~\cite{Avrachenkov2012Top-k}. The algorithm in \cite{Avrachenkov2012Top-k} is a randomized algorithm for undirected graphs that finds a top-$k$ list of nodes with largest degrees in sublinear time. This algorithm is based on a random walk with uniform jumps, described by the following transition probabilities \cite{ART10}: \begin{equation}\label{eq:probrestart} p_{ij} = \left\{ \begin{array}{ll} \frac{\alpha/N+1}{d_i+\alpha}, & \mbox{if $i$ has a link to $j$},\\ \frac{\alpha/N}{d_i+\alpha}, & \mbox{if $i$ does not have a link to $j$}, \end{array}\right. \end{equation} where $N$ is the number of nodes in the graph and $d_i$ is the degree of node $i$. The parameter $\alpha$ controls how often the random walk makes an artificial jump. In \cite{Avrachenkov2012Top-k} it is suggested to take $\alpha$ equal to the average degree in order to maximize the number of independent samples, where the probability of sampling a node is proportional to its degree. After $n$ steps of the random walk, the algorithm returns top-$k$ degree nodes from the set of all visited nodes. See Algorithm~\ref{algo2} for formal description. \IncMargin{1em} \begin{algorithm} \SetKwInOut{Input}{input}\SetKwInOut{Output}{output} \Input{Undirected graph $G$ with $N$ nodes, number of steps $n$, size of output list $k$, parameter $\alpha$} \Output{Nodes $v_1, \dots v_{k}$, their degrees $d_1, \dots, d_{k}$} \BlankLine $v \leftarrow random(N)$\; $A \leftarrow Neighbors(v)$\; $D[v] \leftarrow size(A)$\; \For{$i\leftarrow 2$ \KwTo $n$}{ $r \xleftarrow{\text{sample}} U[0, 1]$\; \If{$r < \frac{D[v]}{D[v]+\alpha}$}{ $v \leftarrow \text{ random from }A$\; } \Else{ $v \leftarrow random(N)$\;} $A \leftarrow Neighbors(v)$\; $D[v] \leftarrow size(A)$\; } $v_1, \dots, v_{k} \leftarrow Top\_k(D)$ // $D[v_1], \ldots, D[v_{k}]$ are the top $k$ maximum values in $D$\; \caption{RandomWalk}\label{algo2} \end{algorithm}\DecMargin{1em} Note that Algorithm~\ref{algo2} works only on undirected graphs. In our implementation on Twitter, all links in the Twitter graph are treated as undirected, and the algorithm returns the top-$k$ in-degree visited vertices. The idea behind this is that the random walk will often find users with large total number of followers plus followees, and since the number of followers of popular users is usually much larger than the number of followees, the most followed users will be found. Another problem of Algorithm~\ref{algo2} in our experimental settings is that it needs to request IDs of all neighbors of a visited node in order to follow a randomly chosen link, while only limited number of IDs can be obtained per one API request (5000 in Twitter). For example, the random walk will quickly find a node with 30M followers, and we will need 6K requests to obtain IDs of all its neighbors. Therefore, an honest implementation of Algorithm~\ref{algo2} usually finds not more than one vertex from top-100. Thus, we have implemented two versions of this algorithm: strict and relaxed. One step of the strict version is one API request, one step of the relaxed version is one considered vertex. Relaxed algorithm runs much longer but shows better results. For both algorithms we took $\alpha = 100$, which is close to twice the average out-degree in Twitter. \textbf{Crawl-Al and Crawl-GAI}~\cite{Kumar2008}. We are given a directed graph $G$ with $N$ nodes. At each step we consider one node and ask for its outgoing edges. At every step all nodes have their \textit{apparent in-degrees} $S_j$, $j=1,\ldots,N$: the number of discovered edges pointing to this node. In Crawl-Al the next node to consider is a random node, chosen with probability proportional to its apparent in-degree. In Crawl-GAI, the next node is the node with the highest apparent in-degree. After $n$ steps we get a list of nodes with largest apparent in-degrees. See Algorithm~\ref{Crawl-GAI} for the pseudocode of Crawl-GAI. \IncMargin{1em} \begin{algorithm} \SetKwInOut{Input}{input}\SetKwInOut{Output}{output} \Input{Directed graph $G$ with $N$ nodes, number of steps $n$, size of output list $k$} \Output{Nodes $v_1, \dots v_{k}$} \BlankLine \For{$i\leftarrow 1$ \KwTo $N$}{ $S[i] \leftarrow 0$\;} \For{$i\leftarrow 1$ \KwTo $n$}{ $v \leftarrow \mathrm{argmax}(S[i])$\; $A \leftarrow OutNeighbors(v)$\; \ForEach{$j$ in $A$}{ $S[j] \leftarrow S[j]+1$\; }} $v_1, \dots, v_{k} \leftarrow Top\_k(S)$ // $S[v_1], \ldots, S[v_{k}]$ are the top $k$ maximum values in $S$\; \caption{Crawl-GAI}\label{Crawl-GAI} \end{algorithm}\DecMargin{1em} \textbf{HighestDegree}~\cite{Brautbar2010high_degree}. A strategy which aims at finding the vertex with largest degree is suggested in \cite{Brautbar2010high_degree}. In our experimental setting with a limited number of API requests this algorithm can be presented as follows. While we have spare resources we choose random vertices one by one and then check the degrees of their neighbors. If the graph is directed, then we check the incoming degrees of out-neighbors of random vertices. See Algorithm~\ref{HighDegree} for the pseudocode of the directed version of this algorithm. \IncMargin{1em} \begin{algorithm} \SetKwInOut{Input}{input}\SetKwInOut{Output}{output} \Input{Directed graph $G$ with $N$ nodes, number of steps $n$, size of output list $k$} \Output{Nodes $v_1, \dots v_{k}$, their degrees $d_1, \dots, d_{k}$} \BlankLine $s \leftarrow 0$\; \For{$i\leftarrow 1$ \KwTo $n$}{\ \If{$s=0$}{ $v \leftarrow random(N)$\; $A \leftarrow OutNeighbors(v)$\; $s \leftarrow size(A)$\; } \Else{ $D[A[s]] \leftarrow InDeg(A[s])$\; $s \leftarrow s-1$\; } } $v_1, \dots, v_{k} \leftarrow Top\_k(D)$ // $D[v_1], \ldots, D[v_{k}]$ are the top $k$ maximum values in $D$\; \caption{HighestDegree}\label{HighDegree} \end{algorithm}\DecMargin{1em} \medskip The algorithms Crawl-AI, Crawl-GAI and HighestDegree find nodes of large in-degrees, but crawl only out-degrees that are usually much smaller. Yet these algorithms can potentially suffer from the API constraints, for example, when in-degrees and out-degrees are positively dependent so that large in-degree nodes tend have high number of out-links to be crawled. In order to avoid this problem on Twitter, we limit the number of considered out-neighbors by 5000 for these algorithms. In the remainder of this section we compare our Algorithm~\ref{algo1} to the baselines on the Twitter follower graph. The first set of results is presented in Table~\ref{tab:comparisonTwitter}, where we take the same budget (number of request to API) $n=1000$ for all tested algorithms to compare their performance. If the standard rate of requests to Twitter API (one per minute) is used, then 1000 requests can be made in 17 hours. For the algorithm suggested in this paper we took $n_1 = 700$, $n_2=300$. As it can be seen from Table~\ref{tab:comparisonTwitter}, Crawl-GAI algorithm, that always follows existing links, seems to get stuck in some densely connected cluster. Note that Crawl-AI, which uses randomization, shows much better results. Both Crawl-GAI and Crawl-AI base their results only on apparent in-degrees. The low precision indicates that due to randomness apparent in-degrees of highest in-degree nodes are often not high enough. Clearly, the weakness of these algorithms is that the actual degrees of the crawled nodes remain unknown. Algorithm~\ref{algo2}, based on a random walk with jumps, uses API requests to retrieve IDs of all neighbors of a visited node, but only uses these IDs to choose randomly the next node to visit. Thus, this algorithm very inefficiently spends the limited budget for API requests. Finally, HighestDegree uses a large number of API requests to check in-degrees of all neighbors of random nodes, so it spends a lot of resources on unpopular entities. Our Algorithm~\ref{algo1} greatly outperforms the baselines. The reason is that it has several important advantages: 1) it is insensitive to correlations between degrees; 2) when we retrieve IDs of the neighbors of a random node (at the first stage of the algorithm), we increase their count of $S$, hence we do not lose any information; 3) sorting by $S[w]$ prevents the waste of resources on checking the degrees of unpopular nodes at the second stage; 4) the second stage of the algorithm returns the exact degrees of nodes, thus, to a large extent, we eliminate the randomness in the values of $S$. \begin{table} \begin{center} \caption{Percentage of correctly identified nodes from top-100 in Twitter averaged over 30 experiments, $n = 1000$}\label{tab:comparisonTwitter} \begin{tabular}{|l|c|c|} \hline Algorithm & mean & standard deviation \\ \hline \hline Two-stage algorithm & 92.6 & 4.7 \\ \hline RandomWalk (strict) & 0.43 & 0.63 \\ \hline RandomWalk (relaxed) & 8.7 & 2.4 \\ \hline Crawl-GAI & 4.1 & 5.9 \\ \hline Crawl-AI & 23.9 & 20.2 \\ \hline HighestDegree & 24.7 & 11.8 \\ \hline \end{tabular} \end{center} \end{table} On Figure~\ref{fig:dynamic_quality} we compare the average performance of our algorithm with the average performance of the baseline strategies for different values of $n$ (from 100 to 5000 API requests). For all values of $n$ our algorithm outperforms other strategies. \begin{figure} \centerline{\includegraphics[width = 0.5\textwidth]{dynamic_quality}} \caption{The fraction of correctly identified top-$100$ most followed Twitter users as a function of $n$ averaged over 10 experiments.} \label{fig:dynamic_quality} \end{figure} \subsection{Finding the largest interest groups}\label{sec:groups} \label{sec:groups} In this section, we demonstrate another application of our algorithm: finding the largest interest groups in online social networks. In some social networks there are millions of interest groups and crawling all of them may not be possible. Using the algorithm proposed in this paper, the most popular groups may be discovered with a very small number of requests to API. In this case, let $V$ be a set of users, $W$ be a set of interest groups, and $(v,w)\in E$ iff $v$ is a member of $w$. Let us demonstrate that our algorithm allows to find the most popular interest groups in the large social network VKontakte with more than 200M registered users. As in the case of Twitter, information on the VKontakte graph can be obtained via API. Again, all users have IDs: integer numbers starting from $1$. Due to this ID assignment, a random user in this network can be easily chosen. In addition, all interest groups also have their own IDs. We are interested in the following requests to API: i) given an ID of a user, return his or her interest groups, ii) given an ID of a group return its number of members. If for some ID there is no user or a user decides to hide his or her list of groups, then an error occurs. The portion of such errors is again approximately $30\%$. As before, first we used our algorithm with $n_1=n_2=50\,000$ in order to obtain the ground truth for the top-100 most popular groups (publicly available sources give the same top-100). Table~\ref{tab:groups} presents some statistics on the most popular groups. \begin{table}[htb] \begin{center} \caption{The most popular groups for VKontakte}\label{tab:groups} \begin{tabular}{|l|c|c|} \hline Rank & Number of participants & Topic \\ \hline \hline 1 & 4,35M & humor \\ \hline 2 & 4,10M & humor \\ \hline 3 & 3,76M & movies \\ \hline 4 & 3,69M & humor \\ \hline 5 & 3,59M & humor \\ \hline 6 & 3,58M & facts \\ \hline 7 & 3,36M & cookery \\ \hline 8 & 3,31M & humor \\ \hline 9 & 3,14M & humor \\ \hline 10 & 3,14M & movies \\ \hline \hline 100 & 1,65M & success stories \\ \hline \end{tabular} \end{center} \end{table} Then, we took $n_1 = 700$, $n_2 = 300$ and computed the fraction of correctly identified groups from top-100. Using only 1000 API requests, our algorithm identifies on average $73.2$ groups from the top-100 interest groups (averaged over 25 experiments). The standard deviation is $4.6$. \section{Performance predictions}\label{sec:analysis} In this section, we evaluate the performance of Algorithm~1 with respect to the metrics (\ref{eq:prediction}) and (\ref{eq:prediction1}) as a function of the algorithm's parameters $n_1$ and $n_2$. Recall that without loss of generality the nodes in $W$ can be numbered $1,2,\ldots, M$ in the decreasing order of their in-degrees, $F_j$ is the unknown in-degree of a node $j$, and $S_j$ is the number of followers of a node $j$ among the randomly chosen $n_1$ nodes in $V$. As prescribed by Algorithm~\ref{algo1}, we pick $n_1$ nodes in $V$ independently and uniformly at random with replacement. If we label all nodes from $V$ that have an edge to $j \in W$, then $S_j$ is exactly the number of labeled nodes in a random sample of $n_1$ nodes, so its distribution is $Binomial\Big(n_1,\frac{F_j}{N}\Big)$. Hence we have \begin{equation} \label{eq:sj_exp} \mathbb{E}(S_j)=n_1\,\frac{F_j}{N},\quad {\rm Var}(S_j)=n_1\,\frac{F_j}{N}\,\Big(1-\frac{F_j}{N}\Big). \end{equation} We are interested in predictions for the metrics (\ref{eq:prediction}) and (\ref{eq:prediction1}). These metrics are completely determined by the probabilities $P_j(n_1)$, $j=1,\ldots,k$, in (\ref{eq:pj}). The expressions for $P_j(n_1)$, $j=1,\ldots,k$, can be written in a closed form, but they are computationally intractable because they involve the order statistics of $S_1,S_2,\ldots,S_M$. Moreover, these expressions depend on the unknown in-degrees $F_1,F_2,\ldots,F_M$. We suggest two predictions for (\ref{eq:prediction}) and (\ref{eq:prediction1}). First, we give a {\it Poisson} prediction that is based on the unrealistic assumption that the degrees $F_1,\ldots,F_{n_2}$ are known, and replaces the resulting expression for (\ref{eq:prediction}) and (\ref{eq:prediction1}) by an alternative expression, which is easy to compute. Next, we {suggest} an {\it Extreme Value Theory} (EVT) prediction that {does not require any preliminary knowledge of unknown degrees but uses the top-$m$ values of highest degrees obtained by the algorithm, where $m$ is much smaller than $k$.} \subsection{Poisson predictions}\label{sec:Poisson} First, for $j=1,\ldots,k$ we write \begin{multline} \label{eq:decompose} P_j(n_1) =\\= \mathrm{P}(S_j>S_{i_{n_2}})+\mathrm{P}(S_j=S_{i_{n_2}},j\in\{i_1,\ldots,i_{n_2}\}). \end{multline} Note that if $[S_j>S_{i_{n_2}}]$ then the node $j$ will be selected by the algorithm, but if $[S_j=S_{i_{n_2}}]$, then this is not guaranteed and even unlikely. This observation is illustrated by the following example. \begin{Example}\label{example:1} Consider the Twitter graph and take $n_1=700$, $n_2=300$. Then the average number of nodes $i$ with $S_i=1$ among the top-$l$ nodes is \[ \sum_{i=1}^{l} \mathrm{P}(\mbox{$S_i=1$}) = \sum_{i=1}^{l} 700\,\frac{F_i}{10^9}\left(1-\frac{F_i}{10^9}\right)^{699}, \] which is $223.3$ for $l=1000$, and it is $19.93$ for $l=n_2=300$. Hence, in this example, we usually see $[S_{i_{300}}= 1]$, however, only a small fraction of nodes with $[S_i=1]$ is selected (on a random basis) into the set $\{i_1,\ldots,i_{300}\}$. \end{Example} Motivated by the above example, we suggest to approximate $P_j(n_1)$ in (\ref{eq:decompose}) by its first term $\mathrm{P}(S_j>S_{i_{n_2}})$. Next, we employ the fact that $S_{n_2}$ has the $n_2$-th highest average value among $S_1,\ldots, S_M$, and we suggest to use $S_{n_2}$ as a proxy for the order statistic $S_{i_{n_2}}$. However, we cannot replace $\mathrm{P}(S_j>S_{i_{n_2}})$ directly by $\mathrm{P}(S_j>S_{n_2})$ because the latter includes the case $[S_j>S_{n_2}=0]$, while with a reasonable choice of parameters it is unlikely to observe $[S_{i_{n_2}}=0]$. This is not negligible as, e.g., in Example~\ref{example:1} we have $\mathrm{P}(S_{n_2}=0)\approx 0.06$. Hence, we propose to approximate $\mathrm{P}(S_j>S_{i_{n_2}})$ by $P(S_j>\max\{S_{{n_2}},1\})$, $j=1,\ldots,n_2$. As the last simplification, we approximate the binomial random variables $S_j$'s by independent Poisson random variables. The Poisson approximation is justified because even for $j = 1, \ldots, k$ the value $F_j/N$ is small enough. For instance, in Example~\ref{example:1} we have $F_1/N\approx 0.04$, so $n_1 F_1/N$ is {$700\cdot 0.04=28$}. Thus, summarizing the above considerations, we propose to replace $P_j(n_1)$ in (\ref{eq:prediction}) and (\ref{eq:prediction1}) by \begin{equation}\label{eq:phat} \hat{P}_j(n_1)=\mathrm{P}(\hat{S}_j>\max\{\hat{S}_{{n_2}},1\}),\; j=1,\ldots,n_2,\end{equation} where $\hat{S}_1,\ldots,\hat{S}_{n_2}$ are independent Poisson random variables with parameters $n_1F_1/N,\ldots, n_1F_{n_2}/N$. We call this method a Poisson prediction for (\ref{eq:prediction}) and (\ref{eq:prediction1}). On Figures~\ref{fig:first_mistake}~and~\ref{fig:prediction}~the results of the Poisson prediction are shown by the green line. We see that these predictions closely follow the experimental results (red line). \subsection{EVT predictions}\label{sec:EVT} \label{eq:evt} {Denote by $\hat F_1>\hat F_2>\cdots>\hat F_k$ the top-$k$ values obtained by the algorithm.} Assume that the actual in-degrees in $W$ are randomly sampled from the distribution $G$ that satisfies (\ref{eq:regular}). Then $F_1>F_2> \cdots>F_M$ are the order statistics of $G$. {The EVT techniques allow to predict high quantiles of $G$ using the top values of $F_i$'s~\cite{Dekkers1989moment_estimator}. However, since the correct values of $F_i$'s are not known, we instead use the obtained top-$m$ values $\hat F_1, \hat F_2, \ldots, \hat F_m$, where $m$ is much smaller than $k$. This is justified for two reasons. First, given $F_j$, $j<k$, the estimate $\hat F_j$ converges to $F_j$ almost surely as $n_1\to\infty$, because, in the limit, the degrees can be ordered correctly using $S_i$'s only according to the strong law of large numbers.} {Second, when $m$ is small, the top-$m$ list can be found with high precision even when $n$ is very modest. For example, as we saw on Figure~\ref{fig:fraction_correct}, we find 50 the most followed Twitter users with very high precision using only 1000 API requests.} {Our goal is to estimate $\hat{P}_j(n_1)$, $j=1,\ldots,k$, using only the values $\hat F_1,\ldots,\hat F_m$, $m<k$. To this end, we suggest to first estimate the value of $\gamma$ using the classical Hill's estimator $\hat{\gamma}$ \cite{Hill} based on the top-$m$ order statistics:} \begin{equation} \label{eq:hill} \hat{\gamma}=\frac{1}{m-1}\sum_{i=1}^{m-1}(\log(\hat F_i)-\log(\hat F_m)). \end{equation} Next, we use the quantile estimator, given by formula (4.3) in \cite{Dekkers1989moment_estimator}, but we replace their two-moment estimator by the Hill's estimator in (\ref{eq:hill}). This is possible because both estimators are consistent (under slightly different conditions). Under the assumption $\gamma>0$, we have the following estimator {$\hat f_j$} for the $(j-1)/M$-th quantile of $G$: \begin{equation} \label{eq:fj} \hat f_j=\hat F_m\left(\frac{m}{j-1}\right)^{\hat{\gamma}},\qquad j>1, j<<M. \end{equation} We propose to use $\hat f_j$ as a prediction of the correct values $F_j$, $j=m+1,\ldots, n_2$. Summarising the above, we suggest the following prediction procedure, which we call EVT prediction. \begin{enumerate} \item Use Algorithm~\ref{algo1} to find the top-$m$ list, $m << k$. \item {Substitute the identified $m$ highest degrees $\hat F_1, \hat F_2,\ldots,\hat F_m$} in (\ref{eq:hill}) and (\ref{eq:fj}) in order to compute, respectively, $\hat\gamma$ and $\hat f_j$, $j=m+1,\ldots,n_2$. \item Use the Poisson prediction (\ref{eq:phat}) substituting the values $F_1,\ldots, F_{n_2}$ by {$\hat F_1,\ldots, \hat F_m$, $\hat f_{m+1},\ldots, \hat f_{n_2}$.} \end{enumerate} On Figures~\ref{fig:prediction}~and~\ref{fig:first_mistake} the blue lines represent the EVT predictions, with $k=100$, $m=20$ and different values of $n_2$. For the average fraction of correctly identified nodes, depicted on Figure~\ref{fig:prediction}, we see that the EVT prediction is very close to the Poisson prediction and the experimental results. The predictions for the first error index on Figure~\ref{fig:first_mistake} are less accurate but the shape of the curve and the optimal value of $n_2$ is captured correctly by both predictors. Note that the EVT prediction tends to underestimate the performance of the algorithm for a large range of parameters. This is because in Twitter the highest degrees are closer to each other than the order statistics of a regularly varying distribution would normally be, which results in an underestimation of $\gamma$ in (\ref{eq:hill}) if only a few top-degrees are used. Note that the estimation \eqref{eq:fj} is inspired but not entirely justified by \cite{Dekkers1989moment_estimator} because the consistency of the proposed quantile estimator (\ref{eq:fj}) is only proved for $j<m$, while we want to use it for $j>m$. However, we see that this estimator agrees well with the data. \begin{figure} \centerline{\includegraphics[width = 0.5\textwidth]{Poisson_Prediction-20}} \caption{Fraction of correctly identified nodes out of top-100 most followed users in Twitter as a function of $n_2$, with $n=1000$.} \label{fig:prediction} \end{figure} \section{Optimal scaling for algorithm parameters} In this section, our goal is to find the ratio $n_2$ to $n_1$ which maximizes the performance of Algorithm~\ref{algo1}. For simplicity, as a performance criterion we consider the expected fraction of correctly identified nodes from the top-$k$ list (see Equation~\eqref{eq:prediction}): $$ \maximize_{n_1, n_2: n_1+n_2=n}\frac{1}{k}\sum_{j=1}^k P_j(n_1)\,. $$ We start with stating the optimal scaling for $n_1$. Let us consider the number of nodes with $S_j>0$ after the first stage of the algorithm. Assuming that the out-degrees of randomly chosen nodes in $V$ are independent, by the strong law of large numbers we have \[\limsup_{n_1\to\infty}\frac{1}{n_1}\sum_{j=1}^MI\{S_j>0\} \le \mu \quad \mbox{ with probability 1}, \] where $\mu$ is the average out-degree in $V$ and $I\{A\}$ is an indicator of the event $A$. Thus, there is no need to check more than $n_2=O(n_1)$ nodes on the second stage, which directly implies the next proposition. \begin{prop} \label{prop:n1} It is optimal to choose $n_1$ such that $n=O(n_1)$. \end{prop} As we noted before (see, e.g., Figure~\ref{fig:fraction_correct}), for small $k$ the algorithm has a high precision in a large range of parameters. However, for not too small values of $k$, the optimization becomes important. In particular, we want to maximize the value $P_k(n_1)$. We prove the following theorem. \begin{theorem} \label{th:n2} Assume that $k=o(n)$ as $n \to \infty$, then the maximizer of the probability $P_k(n-n_2)$ is $$ n_2= \left( 3 \gamma k^{\gamma} n \right)^{\frac{1}{\gamma+1}}\left(1 + \mbox{o}(1)\right), $$ with $\gamma$ as in $(\ref{eq:regular})$. \end{theorem} \begin{proof} It follows from Proposition~\ref{prop:n1} that $n_1 \to \infty$ as $n \to \infty$, so we can apply the following normal approximation \begin{align} \nonumber P_k(n_1)& \approx P\left(N\left(\frac{n_1(F_k-F_{n_2})}{N},\frac{n_1(F_k+F_{n_2})}{N}\right)>0\right)\\ \label{eq:normal} &= P\left(N(0,1) > -\sqrt{\frac{n_1}{N}}\frac{F_k - F_{n_2}}{\sqrt{F_k+F_{n_2}}}\right). \end{align} The validity of the normal approximation follows from the Berry-Esseen theorem. In order to maximize the above probability, we need to maximize $\sqrt{\frac{n_1}{N}}\frac{F_k - F_{n_2}}{\sqrt{F_k+F_{n_2}}}$. It follows from EVT that $F_k$ decays as $k^{-\gamma}$. So, we can maximize \begin{align}\label{eq:optim} \frac{\sqrt{n-n_2}\left(k^{-\gamma} - n_2^{-\gamma}\right)}{\sqrt{k^{-\gamma}+n_2^{-\gamma}}}. \end{align} Now if $n_2=\mbox{O}(k)$, then $\sqrt{n-n_2}=\sqrt{n}(1+o(1))$ and the maximization of (\ref{eq:optim}) mainly depends on the remaining term in the product, which is an increasing function of $n_2$. This suggests that $n_2$ has to be chosen considerably greater than $k$. Also note that it is optimal to choose $n_2=o(n)$ since only in this case the main term in \eqref{eq:optim} amounts to $\sqrt{n}$. Hence, we proceed assuming the only interesting asymptotic regime where $k=o(n_2)$ and $n_2=o(n)$. In this asymptotic regime, we can simplify (\ref{eq:optim}) as follows: $$ \frac{\sqrt{n-n_2}\left(k^{-\gamma} - n_2^{-\gamma}\right)}{\sqrt{k^{-\gamma}+n_2^{-\gamma}}}= $$ $$ \frac{1}{k^{\gamma/2}} \sqrt{n-n_2} \left( 1- \frac{3}{2} \left(\frac{k}{n_2}\right)^\gamma + \mbox{O}\left(\left(\frac{k}{n_2}\right)^{2\gamma}\right)\right). $$ Next, we differentiate the function $$ f(n_2):= \sqrt{n-n_2} \left( 1- \frac{3}{2} \left(\frac{k}{n_2}\right)^\gamma\right) $$ and set the derivative to zero. This results in the following equation: \begin{equation}\label{eq:n2eq} \frac{1}{3\gamma k^\gamma} n_2^{\gamma+1} + n_2 - n = 0. \end{equation} Since $n_2 = \mbox{o}(n)$, then only the highest order term remains in (\ref{eq:n2eq}) and we immediately obtain the following approximation $$ n_2= \left( 3 \gamma k^{\gamma} n \right)^{\frac{1}{\gamma+1}}\left(1 + \mbox{o}(1)\right). $$ \end{proof} \section{Sublinear complexity}\label{sec:complexity} The normal approximation (\ref{eq:normal}) implies the following proposition. \begin{prop} \label{prop:1} For large enough $n_1$, the inequality \begin{equation} \label{eq:prop1} Z_k(n_1):=\sqrt{\frac{n_1}{N}}\frac{F_k - F_{n_2}}{\sqrt{F_k+F_{n_2}}} \ge z_{1-\varepsilon}, \end{equation} where $z_{1-\varepsilon}$ is the $(1-\varepsilon)$-quantile of a standard normal distribution, guarantees that the mean fraction of top-$k$ nodes in $W$ identified by Algorithm~\ref{algo1} is at least $1-\varepsilon$. \end{prop} Using (\ref{eq:fj}), the estimated lower bound for $n_1$ in (\ref{eq:prop1}) is: \begin{equation} \label{eq:rough} n_1 \ge \frac{N z_{1-\varepsilon}^2 (k^{-\hat{\gamma}}+n_2^{-\hat{\gamma}})}{\hat F_m m^{\hat \gamma}(k^{-\hat{\gamma}}-n_2^{-\hat{\gamma}})^2}. \end{equation} In the case of the Twitter graph with $N=10^9$, $m=20$, $\hat F_{20}=18,825,829$, $k=100$, $n_2=300$, $z_{0.9}\approx 1.28$, $\hat\gamma=0.4510$, this will result in $n_1\ge 1302$, which is more pessimistic than $n_1=700$ but is sufficiently close to reality. Note that Proposition~\ref{prop:1} is expected to provide a pessimistic estimator for $n_1$, since it uses the $k$-th highest degree, which is much smaller than, e.g., the first or the second highest degree. We will now express the complexity of our algorithm in terms of $M$ and $N$, assuming that the degrees in $W$ follow a regularly varying distribution $G$ defined in (\ref{eq:regular}). In a special case, when our goal is to find the highest in-degree nodes in a directed graph, we have $N=M$. If $M$ is, e.g., the number of interest groups, then it is natural to assume that $M$ scales with $N$ and $M\to\infty$ as $N\to\infty$. Our results specify the role of $N$, $M$, and $G$ in the complexity of Algorithm~\ref{algo1}. From (\ref{eq:rough}) we can already anticipate that $n$ is of the order smaller than $N$ because $F_m$ grows with $M$. This argument is formalized in Theorem~\ref{th:complexity} below. \begin{theorem} \label{th:complexity} Let the in-degrees of the entities in $W$ be independent realizations of a regularly varying distribution $G$ with exponent $1/\gamma$ as defined in (\ref{eq:regular}), and $F_1\ge F_2\ge\cdots\ge F_M$ be their order statistics. Then for any fixed $\varepsilon,\delta>0$, Algorithm~\ref{algo1} finds the fraction $1-\varepsilon$ of top-$k$ nodes with probability $1-\delta$ in \[n=\mbox{O}(N/a(M))\] API requests, as $M,N\to\infty$, where $a(M)=l(M)M^\gamma$ and $l(\cdot)$ is some slowly varying function. \end{theorem} \begin{proof} Let $a(\cdot)$ be a left-continuous inverse function of $1/(1-G(x))$. Then $a(\cdot)$ is a regularly varying function with index $\gamma$ (see, e.g., \cite{BinGolTeu89}), that is, $a(y)=l(y)y^{\gamma}$ for some slowly varying function $l(\cdot)$. Furthermore, repeating verbatim the proof of Theorem~2.1.1 in \cite{deHaan-Ferreira}, we obtain that for a fixed $m$ \[\left(\frac{F_1}{a(M)},\cdots,\frac{F_m}{a(M)}\right)\stackrel{d}{\to}\left(E_1^{-\gamma},\cdots,(E_1+\cdots+E_m)^{-\gamma}\right),\] where $E_i$ are independent exponential random variables with mean~1 and $\stackrel{d}{\to}$ denotes the convergence in distribution. Now for fixed $k$, choose $n_2$ as in Theorem~\ref{th:n2}. It follows that if $n_1=CN/a(M)$ for some constant $C>0$ then $Z_k(n_1)\stackrel{d}{\to}\sqrt{C(E_1+\cdots+E_k)^{-\gamma}}$ as $M,N\to\infty$. Hence, we can choose $C$, $M$, $N$ large enough so that $P(Z_k(n_1)>z_{1-\varepsilon})>1-\delta$. We conclude that $n_1=\mbox{O}(N/a(M))$ for fixed $k$, as $N,M\to\infty$. Together with Proposition~\ref{prop:n1}, this gives the result. \end{proof} In the case $M=N$, as in our experiments on Twitter, Theorem~\ref{th:complexity} states that the complexity of the algorithm is roughly of the order $N^{1-\gamma}$, which is much smaller than linear in realistic networks, where we often observe $\gamma\in(0.3,1)$~\cite{Newman}. The slowly varying term $l(N)$ does not have much effect since it grows slower than any power of $N$. In particular, if $G$ is a pure Pareto distribution, $1-G(x)=Cx^{-1/\gamma}$, $x\ge x_0$, then $a(N)=C^{\gamma} N^{\gamma}$. \section{Conclusion}\label{sec:conclusion} In this paper, we proposed a randomized algorithm for quick detection of popular entities in large online social networks whose architecture has underlying directed graphs. Examples of social network entities are users, interest groups, user categories, etc. We analyzed the algorithm with respect to two performance criteria and compared it with several baseline methods. Our analysis demonstrates that the algorithm has sublinear complexity on networks with heavy-tailed in-degree distribution and that the performance of the algorithm is robust with respect to the values of its few parameters. Our algorithm significantly outperforms the baseline methods and has much wider applicability. An important ingredient of our theoretical analysis is the substantial use of the extreme value theory. The extreme value theory is not so widely used in computer science and sociology but appears to be a very useful tool in the analysis of social networks. We feel that our work could provide a good motivation for wider applications of EVT in social network analysis. We validated our theoretical results on two very large online social networks by detecting the most popular users and interest groups. We see several extensions of the present work. A top list of popular entities is just one type of properties of social networks. We expect that both our theoretical analysis, which is based on the extreme value theory, and our two-stage randomized algorithm can be extended to infer and to analyze other properties such as the power law index and the tail, network functions and network motifs, degree-degree correlations, etc. It would be very interesting and useful to develop quick and effective statistical tests to check for the network assortativity and the presence of heavy tails. Since our approach requires very small number of API requests, we believe that it can be used for tracing network changes. Of course, we need formal and empirical justifications of the algorithm applicability for dynamic networks. \section*{Acknowledgment} This work is partially supported by the EU-FET Open grant NADINE (288956) and CONGAS EU project FP7-ICT-2011-8-317672. \bibliographystyle{abbrv}
\section{Introduction} The question of how to understand the discrepancy between two samples is of fundamental importance in various disciplines such as statistics, information theory and machine learning. In statistics, hypothesis testing procedures such as Kolmogorov-Simirnov test and Cramer-von Mises test are based on some discrepancy measurement of the empirical distributions, and a common goal in Bayesian experimental design \citep{Chaloner1995} is to maximize the expected Kullback-Leibler (KL) divergence between the prior and the posterior. In information theory, divergence plays an important role since other information theoretic quantities such as mutual information and Shannon Entropy are derived from it. In machine learning, probability distance measures such as Total Variation, Hellinger Distance and KL divergence are often applied as a loss function in manifold learning \citep{Donoho2003}, classification and anomaly detection, etc. For instance, it is applied to image registration and multimedia classification problems as a similarity measure \citep{Ho2004}. Estimating discrepancies of two probability distributions, especially KL divergence, based on a set of observations from each distribution has a long history and rich literature. For ease of elaboration, we formulate the problem in mathematical terms: given the sample space $\Omega$, two sets of i.i.d observations $\mathcal{X} = \{x_i\}_{i = 1}^{n_1}$ and $\mathcal{Y} = \{y_i\}_{i = 1}^{n_2}$ are generated from probability measures $P_1$ and $P_2$ with densities $p_1(x)$ and $p_2(y)$ respectively. Various measures of discrepancy of $p_1$ and $p_2$ have the following general form \[D_\phi(p_1, p_2) = \int p_1(x)\phi\Big(\frac{p_2(x)}{p_1(x)}\Big)dx\] where $\phi$ is a convex function. $D_\phi$ is the class of Ali-Silvey distances \citep{Ali1966}, also known as $f-$divergences. Many common divergences, such as KL divergence ($\phi(x) = -\log(x)$), $\alpha-$divergence (e.g., $\phi(x) = 4 / (1 - \alpha^2)\cdot(1 - x^{(\alpha + 2) / 2})$ when $\alpha \neq \pm 1$), Hellinger distance ($\sqrt{D_\phi}$ and $\phi(x) = 1 - \sqrt{x}$), and total variation distance ($\phi(x) = |x - 1|$), are special cases of $f-$divergence, coinciding with a particular choice of $f$. We are interested in estimating $D_\phi$ from two samples $\mathcal{X}$ and $\mathcal{Y}$, without any assumptions of the forms of $p_1$ and $p_2$. A conceptually simple approach is to proceed in two-steps: estimate the densities $\hat{p}_1$ and $\hat{p}_2$ according to the $\mathcal{X}$ and $\mathcal{Y}$ independently then calculate $D_\phi(\hat{p}_1, \hat{p}_2)$. However, this method is unattractive for the following reasons: i) multivariate density estimation itself is a challenging problem and often more difficult than comparing distributions, so poorly estimated densities can cause subsequent discrepancy estimation to suffer from high variance and bias \citep{Ma2011}; ii) if the first stage is carried out independently, there is no easy way to compute their discrepancy $D_\phi(\hat{p}_1, \hat{p}_2)$ analytically; some numeric or approximation routines have to be employed, which in turn introduce more error terms \citep{Sugiyama2012}. Hence, throughout this paper, we focus on the class of single-shot methods, i.e., comparing samples directly and simultaneously. We review several methods in the context of information theory and machine learning. Interested readers may refer to \citep{Sricharan2010, Wang2009, Leonenko2008, Poczos2011, Ma2011, Poczos2012} for other recent developments. \citep{Wang2005} proposes a domain-splitting method that constructs an adaptive partition of the sample space with respect to $\mathcal{Y}$, then Radon-Nikodym derivative $p_1(x) / p_2(x)$ can be estimated with empirical probability mass (with the correction term on the boundary) in each sub-region. In 1-dimension, strong consistency is established and several algorithmic refinements are suggested. In multi-dimensions, a partitioning scheme is also mentioned based on the heuristics in 1-dimensional case, however, there are not enough numeric simulations to justify the heuristics. \citep{Nguyen2010} derives a variational characterization of the $f-$divergence in terms of a Bayes decision problem, which is exploited to develop an $M-$estimator. The theoretical results of consistency and convergence rates are provided. It is also shown that the estimation procedure can be cast into a finite-dimensional convex program and solved efficiently when the function class is defined by reproducing kernel Hilbert spaces. \citep{Sugiyama2012} introduces the least-squares density-difference (LSDD) estimator is developed in the framework of kernel regularized least-squares estimation. The finite sample error bound shows that LSDD achieves the optimal convergence rate. They also demonstrate several pattern recognition applications. Despite of the success of these methods, they are insufficient in many scenarios. For example, since $\mathcal{X}$ and $\mathcal{Y}$ are two sets of random observations, one may be interested in the confidence (credible) intervals of estimated discrepancies \citep{Sricharan2010}, but deriving such quantities are non-trivial for these methods. Secondly, these methods and theories are all set up for a specific type of discrepancy and are usually not directly applicable to others. Thirdly, to the best of our knowledge, few experiments are conducted in dimensions larger than 3 in previous work \citep{Sricharan2010, Nguyen2010, Perez-Cruz2008}, even though most of the methods claim that there are no technical difficulties as one increases the number of dimensions. Finally, there is little work on tackling discrepancy estimation under the Bayesian perspective. In this paper, we introduce a single-shot Bayesian model, which we name as {\bf co}upled {\bf B}inary {\bf P}artition {\bf M}odel (co-BPM). Our model design is inspired by two key ideas. First, we use an adaptive partitioning scheme, which has been demonstrated to be more scalable (see Table \ref{extreme}) than traditional domain splitting methods such as histogram~\citep{Wang2005} or regular paving~\citep{Sainudiin2013}. Second, instead of partitioning based upon $\mathcal{Y}$ alone as in~\citep{Wang2005}, we force the domain $\Omega$ to be \emph{coupled} so that it is partitioned with respect to $\mathcal{X}$ and $\mathcal{Y}$ simultaneously. Therefore, our model is capable of capturing the landscape of both $\mathcal{X}$ and $\mathcal{Y}$, thus is more effective in estimating their discrepancies. We highlight our contributions as follows: \begin{enumerate} \item To our knowledge, co-BPM is the first \emph{single-shot domain-partition} based Bayesian model for discrepancy estimation, that is scalable in both dimension and sample size. \item The specifically tailored MCMC sampling algorithm exploits the sequential buildup of binary partition for rapid mixing. \item The method is tested in 5 and 10 dimensional settings that are rare in previous work. Experiments show that we achieve better convergence and higher accuracy than other methods. \item We propose two \emph{variance reduction} approaches that are effective for discrepancies involving divisions (e.g., in $\alpha-$divergence). \end{enumerate} We validate our model by intensive simulations: 1) two sanity tests are used to demonstrate that our model is sensitive to the differences among samples; 2) a 4-dim example is used to demonstrate the effectiveness of variance reduction; 3) examples in 5-dim and 10-dim are carried out and are compared to the methods in \citep{Nguyen2010} and \citep{Perez-Cruz2008} The rest of the paper is organized as follows: Section 2 introduces some notations used throughout our discussion and preliminaries that motivate our model; Section 3 introduces the Coupled Binary Partition Model and the specially tailored MCMC algorithm to sample posterior effectively; extensive simulations and comparisons are presented in Section 4; we draw the conclusions in Section 5. \begin{figure} \center \includegraphics[width = 0.8\textwidth]{sbp.pdf} \caption{A \textbf{binary partition sequence} demonstrates its sequential structure. Each red line indicates the decision to halve the sub-rectangle alone it.} \label{fig:sbp} \end{figure} \begin{figure} \center \includegraphics[width = .4\textwidth]{joint_par.pdf} \caption{An illustration of \textbf{coupled binary partition} of two samples through the coupled binary partition model. Two samples are drawn from two different distributions and their sampling spaces are partitioned in tandem.} \label{joint_par} \end{figure} \section{Notation and Preliminaries} Throughout the paper, we assume $\Omega = [0, 1]^d$ and consider the domain-splitting approach to estimate the discrepancies through coordinate-wise binary partition. A binary partition has hierarchical structure that is constructed sequentially: starting with $\mathcal{B}_1 = \{r_{1, 1} = \Omega\}$ at depth $1$, action $a_1$ is taken to split $\Omega$ into $\mathcal{B}_2 = \{r_{1, 2}, r_{2, 2}\}$ along the middle of one of its coordinates; then at depth 2, action $a_2$ is taken to split one of sub-rectangles in $B_2$ into $\mathcal{B}_3 = \{r_{1, 3}, r_{2, 3}, r_{3, 3}\}$ evenly. The process keeps on till the specified depth is reached. Given the maximum depth $l$, the decision sequence is denoted as $A_l = (a_1, ..., a_{l - 1})$. $A_l$ uniquely determines a split of $\Omega$, which is denoted as $\{r_{1, l}, ..., r_{l, l}\}$. It is necessary to point out that multiple decision sequences may result in the same partition. An important property of binary partition motivates our coupling scheme in the following section. Figure \ref{fig:sbp} gives an illustration of the sequence in $[0, 1]^2$; Figure \ref{joint_par} demonstrates the coupled binary partition, where the domains of two different samples are partitioned in tandem. \begin{pro} For any pair of decision sequences $A_{l_1}$ and $A_{l_2}$ and their partitions $\{r_{i, l_1}\}_{i = 1}^{l_1}$ and $\{r_{i, l_2}\}_{i = 1}^{l_2}$, there exists a sequence $A_l$ and its partition $\{r_{i, l}\}_{i = 1}^l$ such that $r_{i, l_k}$ is the union of a subset of $\{r_{i, l}\}_{i = 1}^l$ for $k = 1, 2$ and $i = 1, ..., l_k$, namely, $A_l$ defines a more refined partition than $A_{l_1}$ or $A_{l_2}$. \label{pro} \end{pro} The next theorem shows that a partition of $\Omega$ gives a lower bound of the discrepancies. \begin{thm} Given a partition $\{r_{1, l}, ..., r_{l, l}\}$ of $\Omega$ at depth $l$, let \[\tilde{p}_i(x) = \sum_{k = 1}^l \frac{P_i(r_{k, l})}{|r_{k, l}|}\mathbf{1}\{x\in r_{k, l}\}\] where $i = 1, 2$ and $|\cdot|$ denotes the volume or size, then \begin{equation} D_\phi(p_1, p_2) \geq D_\phi(\tilde{p}_1, \tilde{p}_2) \label{lower_bound} \end{equation} namely, a partition gives a way to estimate $D_\phi$. \label{thm1} \end{thm} \begin{rem} Another lower bound \citep{Nguyen2010} of $D_\phi$ is $\sup_{f\in\mathcal{F}}\int [fdP_2 - \phi^*(f)dP_1]$, where $\mathcal{F}$ is a class of functions and $\phi^*$ is the conjugate dual function of $\phi$. Theorem \ref{thm1} shifts the difficulty of estimation from finding a good $\mathcal{F}$ to a good partition. \end{rem} \begin{rem} Under the condition that the integral $D_\phi(p_1, p_2)$ is Riemann integrable. It is trivial to show that $D_\phi(p_1, p_2) = \sup D_\phi(\tilde{p}_1, \tilde{p}_2)$, where the supremum is taken over all possible partitions on all depths. \label{rie} \end{rem} The proof is straightforward by applying the Jensen's inequality. {\small \begin{proof} We decompose the integral by $\{r_{1, l}, ..., r_{l, l}\}$ and apply Jensen's inequality, \begin{equation} \begin{split} &D_\phi(p_1, p_2) = \int_{\Omega} p_1(x)\phi\Big(\frac{p_2(x)}{p_1(x)}\Big)dx = \sum_{i = 1}^l\int_{r_{i, l}}p_1(x)\phi\Big(\frac{p_2(x)}{p_1(x)}\Big)dx\\ &= \sum_{i = 1}^l P_1(r_{i, l})E_{p_1(\cdot|r_{i, l})}[\phi\Big(\frac{p_2(x)}{p_1(x)}\Big)] \geq \sum_{i = 1}^l P_1(r_{i, l})\phi\Big(E_{p_1(\cdot|r_{i, l})}\frac{p_2(x)}{p_1(x)}\Big) = \sum_{i = 1}^l P_1(r_{i, l})\phi\Big(\frac{P_2(r_{i, l})}{P_1(r_{i, l})}\Big) \end{split} \end{equation} \end{proof} } Remark \ref{rie} indicates that the divergence can be defined over countable partitions. A simple illustration in Table \ref{extreme} implies the importance of a good partition. In next section, we describe how Theorem \ref{thm1} specializes to a partition based estimator for $f-$divergence. A closer look into the proof of Theorem \ref{thm1} gives us more insights. Let's take the derivation for $D_\phi$ as an example. Within each region, the gap between $D_\phi(p_1, p_2)$ and $D_\phi(\tilde{p}_1, \tilde{p}_2)$ comes from applying Jensen's inequality, therefore, it can be closed if $p_1(x)/p_2(x)$ in each region is approximately constant. Such observations indicate that partitioning the domain more finely would reduce the estimation bias. However, an overly fine partitioning will cause insufficient samples in each region and inadvertently increase the overall estimation variance. So there is a \emph{trade-off between bias and variance}. Therefore, an appropriate partitioning respecting the tradeoff should reflect the landscape of $\mathcal{X}$ and $\mathcal{Y}$ simultaneously and avoid over-cutting. We will see how this intuition is implemented in our Bayesian model. \begin{table} \begin{tabular}{cc} \putindeepbox[10pt]{\includegraphics[width = .4\textwidth, height = .63in]{joint_par_sim_extre.pdf}} & \putindeepbox[10pt]{\begin{tabular}{rcccc} \hline & $D_{q = 1}$ & $D_{q = 2}$ & $D_{\alpha = 1}$ & $D_{\alpha = 2}$\\ \hline Hist & 0.5269 & 0.4935 & 0.7964 & 1.0291 \\ BP & \textbf{0.5431} & \textbf{0.4990} & \textbf{0.8267} & \textbf{1.0527} \\ Truth & 0.5518 & 0.5204 & 0.8604 & 1.0769 \\ \hline \end{tabular}} \end{tabular} \caption{\textbf{A simple illustration demonstrates the importance of a good partition.} (I) $\mathcal{B}_1 = \{\Omega_{11}, \Omega_{12}\}$ and $\mathcal{B}_2 = \{\Omega_{21}, \Omega_{22}\}$ are two partitions of unit cube. Consider $p_1(x) = \frac{3}{2}\mathbf{1}\{x\in \Omega_{11}\} + \frac{1}{2}\mathbf{1}\{x\in \Omega_{12}\}$ and $p_2(x) = \frac{1}{2}\mathbf{1}\{x\in \Omega_{11}\} + \frac{3}{2}\mathbf{1}\{x\in \Omega_{12}\}$. Under $\mathcal{B}_1$, the lower bounds equal $D_\phi(p_1, p_2)$, i.e., $D_\phi = \frac{3}{4}\phi(\frac{1}{3}) + \frac{1}{4}\phi(3)$. For instance, the Total Variation Distance is $\frac{1}{2}$, Hellinger Distance is $\frac{\sqrt{3}}{2} - \frac{1}{2}$ and $\alpha-$divergence is $\alpha = 2$ is $\frac{1}{\alpha - 1}\log(\frac{3}{4}\cdot 3^{\alpha - 1} + \frac{1}{4}\cdot \frac{1}{3}^{\alpha - 1})$ respectively; however, under $\mathcal{B}_2$, the lower bounds are $\phi(1)$, where the Total Variation Distance, Hellinger Distance and $\alpha-$divergence with $\alpha = 2$ are all 0. (II) $p_1(x) = \beta_{3, 5}(x_1)\beta_{3, 5}(x_2)$ and $p_2(x) = \mathbf{1}\{x\in [0, 1]^2\}$. On the left is histogram by dividing each dimension into 8 equal sub-intervals; on the right is an adaptive partition with 64 sub-rectangles. As summarized in the right table, \textbf{BP} (adaptive binary partition) is closer to \textbf{Truth} (the true values) than \textbf{Hist} (histogram).} \label{extreme} \end{table} \section{Coupled Binary Partition Model} In light of Theorem \ref{thm1} and the importance of partitions, we model probability mass pairs $(P_1(r_{i, l}), P_2(r_{i, l}))_{i = 1}^l$ directly for a given partition $\{r_{i, l}\}_{i = 1}^l$. Since the lower bounds \eqref{lower_bound} need only $(P_1(r_{i, l}), P_2(r_{i, l}))_{i = 1}^l$, once they are fixed, we restrict the density in each sub-rectangle to be constant to reduce the complexity of density class, i.e., the density class is piecewise constant functions supported on binary partitions. Specifically, the prior is constructed as follows: 1) the decision sequence of depth $l\in\mathcal{N}^+$ has a prior density proportional to $\exp(-\sigma l)$ with some positive $\sigma$ and that all sequences at the same depth are distributed uniformly---this part of the prior is to discourage over-cutting as discussed in the previous section; 2) given an decision sequence $A_l = (a_1, ..., a_{l - 1})$ and the fact that $\sum_{i = 1}^l P_k(r_{i, l}) = 1$ for $k = 1, 2$, the probability masses $(P_1(r_{i, l}))_{i = 1}^l, (P_2(r_{i, l}))_{i = 1}^l$, denoted as $\mathbf{m}_{1, l}, \mathbf{m}_{2, l}$ or $(m_{1i, l})_{i = 1}^l, (m_{2i, l})_{i = 1}^l$ have a Dirichlet prior $\textrm{Dir}(\delta, ..., \delta)$ respectively with some positive $\delta$ and the priors are independent between samples $\mathcal{X}$ and $\mathcal{Y}$. Thus, \normalsize\[p(l, A_l, (\mathbf{m}_{1, l}, \mathbf{m}_{2, l}))\propto \exp(-\sigma l)\prod_{i = 1}^l m_{1i, l}^{\delta - 1}\prod_{i = 1}^l m_{2i, l}^{\delta - 1}\]\normalsize The densities are uniform conditioned on each sub-rectangle, i.e., \normalsize\[\hat{p}_1(x) = \sum_{i = 1}^l \frac{m_{1i, l}}{|r_{i, l}|}\mathbf{1}\{x\in r_{i, l}\}, \hat{p}_2(y) = \sum_{i = 1}^l \frac{m_{2i, l}}{|r_{i, l}|}\mathbf{1}\{y\in r_{i, l}\}\]\normalsize As discussed in Section 1, a naive two step algorithm proceeds as follows: estimating two piecewise constant densities $\hat{p}_1$ and $\hat{p}_2$ supported on (different) binary partitions independently based on $\mathcal{X}$ and $\mathcal{Y}$; then computing $D_\phi(\hat{p}_1, \hat{p}_2)$, which requires to consider all the intersections of sub-rectangles between two partitions. The drawbacks of this type of approach are already discussed. However, according to Property \ref{pro}, any two piecewise constant densities supported on different binary partitions can be rewritten such that they are defined on the same binary partition, thus we \emph{couple} $\hat{p}_1, \hat{p}_2$ by partitioning their domain in tandem (i.e., forcing the same $A_l$) in the prior such that Theorem \ref{thm1} is applicable. The difference of $\hat{p}_1, \hat{p}_2$ is captured by $\mathbf{m}_{1, l}, \mathbf{m}_{2, l}$. The likelihood of $\mathcal{X}$ and $\mathcal{Y}$ is \normalsize\[p(\mathcal{X}, \mathcal{Y}|(l, A_l, (\mathbf{m}_{1, l}, \mathbf{m}_{2, l}))) = \prod_{i = 1}^l\Big(\frac{m_{1i, l}}{|r_{i, l}|}\Big)^{n_{1i, l}}\prod_{i = 1}^l\Big(\frac{m_{2i, l}}{|r_{i, l}|}\Big)^{n_{2i, l}}\]\normalsize where $(n_{ki, l})_{i = 1}^l, k = 1, 2$ denotes the number of points in each region for samples $\mathcal{X}, \mathcal{Y}$. Under our assumption, the prior of $(\mathbf{m}_{1, l}, \mathbf{m}_{2, l})$ is distributed with joint Dirichlet distribution $\textrm{Dir}(\delta, ..., \delta)\times\textrm{Dir}(\delta, ..., \delta)$. By denoting $\Theta = (l, A_l, (\mathbf{m}_{1, l}, \mathbf{m}_{2, l}))$, the posterior is \normalsize \begin{equation} \begin{split} & p(\Theta = (l, A_l, (\mathbf{m}_{1, l}, \mathbf{m}_{2, l}))|\mathcal{X}, \mathcal{Y})\propto p(\Theta)p(\mathcal{X}, \mathcal{Y}|\Theta)\\ & \propto\exp(-\sigma l)\prod_{i = 1}^l\Big(\frac{1}{|r_{i, l}|}\Big)^{n_{1i, l} + n_{2i, l}}\prod_{i = 1}^l (m_{1i, l})^{\delta + n_{1i, l} - 1}\prod_{i = 1}^l (m_{2i, l})^{\delta + n_{2i, l} - 1} := \psi(\Theta|\mathcal{X}, \mathcal{Y}) \end{split} \label{post} \end{equation} \normalsize Through sampling from the posterior, the discrepancies are estimated by $D_\phi(\hat{p}_1, \hat{p}_2)$. \subsection{Sampling} According to the decomposition in \eqref{post}, the probability masses and the decision sequence are sampled hierarchically. The first step is to generate the depth $l$ and decisions $A_l$. By marginalizing $\mathbf{m}_{1, l}$ and $\mathbf{m}_{2, l}$ in \eqref{post}, $(A_l, l)$ is distributed as \normalsize \begin{equation} p(A_l, l|\mathcal{X}, \mathcal{Y})\propto \exp(-\sigma l)\beta((\delta + n_{1i, l})_{i = 1}^l)\beta((\delta + n_{2i, l})_{i = 1}^l)\prod_{i = 1}^l\Big(\frac{1}{|r_{i, l}|}\Big)^{n_{1i, l} + n_{2i, l}} \label{post_par} \end{equation} \normalsize where $\beta(\cdot)$ is the multinomial Beta function and defined as $\beta((z_i)_{i = 1}^k) = \prod_{i = 1}^k\Gamma(z_i) / \Gamma(\sum_{i = 1}^k z_i)$ and $\Gamma(z)$ is the Gamma function. Furthermore, conditioned on $l$, $A_l$ is distributed as \normalsize \begin{equation} p(A_l|l, \mathcal{X}, \mathcal{Y})\propto \beta((\delta + n_{1i, l})_{i = 1}^l)\beta((\delta + n_{2i, l})_{i = 1}^l)\prod_{i = 1}^l\Big(\frac{1}{|r_{i, l}|}\Big)^{n_{1i, l} + n_{2i, l}} \label{post_par} \end{equation} \normalsize once $A_l$ is generated, $\mathbf{m}_{1, l}$ and $\mathbf{m}_{2, l}$ are sampled through $\textrm{Dir}((\delta + n_{1i, l})_{i = 1}^l)$ and $\textrm{Dir}((\delta + n_{2i, l})_{i = 1}^l)$ respectively. It is difficult to obtain the analytical distribution of $(A_l, l)$ because of the intractability of normalizing constant or partition function, Markov Chain Monte Carlo is employed to sample the posterior \eqref{post}. However, given the vast parameter space and the countless local modes, the naive Metropolis-Hastings \citep{hastings1970} suffers from slow mixing in our experience. In order to sample effectively, a proposal kernel should be chosen with the two properties: 1) it leverages the sequential structure of binary partition and the proposed partition is generated with respect to $\mathcal{X}$ and $\mathcal{Y}$ for rapid mixing; 2) the corresponding acceptance ratio depends on a smaller set of parameters; in other words, the transition probability is controlled by a subset of $\Theta$, such that the size (dimensionality) of searching space is reduced. Define \normalsize \begin{equation*} g(\Theta' = (l', A_{l'}, (\mathbf{m}_{1, l'}, \mathbf{m}_{2, l'}))|\Theta = (l, A_l, (\mathbf{m}_{1, l}, \mathbf{m}_{2, l})))=p(A_{l'}, l'|A_l, l)p((\mathbf{m}_{1, l'}, \mathbf{m}_{2, l'})|A_{l'}, l) \end{equation*} \normalsize where $p(A_{l'}, l'|A_l, l) = p(l'|l)p(A_{l'}|A_l, l')$ defines the jump probability and $l'$ is constrained to be $l - 1$ and $l + 1$. In order to exploit the sequential structure of binary partition such that each $a_l$ is drawn with the guidance of $A_l$, $p(A_{l + 1}|A_l, l + 1) = p(a_l|A_l, l + 1)$ is defined as \begin{equation} \begin{split} p(a_l|A_l, l + 1)& = \frac{p(A_{l + 1}|l + 1, \mathcal{X}, \mathcal{Y})}{p(A_{l}|l, \mathcal{X}, \mathcal{Y})}\\ &\propto\frac{\beta((\delta + n_{1i, l + 1})_{i = 1}^{l + 1})\beta((\delta + n_{2i, l + 1})_{i = 1}^{l + 1})}{\beta((\delta + n_{1i, l})_{i = 1}^l)\beta((\delta + n_{2i, l})_{i = 1}^l)}\frac{\prod_{i = 1}^l |r_{i, l}|^{n_{1i, l} + n_{2i, l}}}{\prod_{i = 1}^{l + 1}|r_{i, l + 1}|^{n_{1i, l + 1} + n_{2i, l + 1}}} \label{individual} \end{split} \end{equation} and $p(A_{l - 1}|A_l, l - 1) = p(l - 1|l)$ since $p(A_{l - 1}|A_l, l - 1) = 1$; $p((\mathbf{m}_{1, l'}, \mathbf{m}_{2, l'})|A_{l'}, l')$ is the joint Dirichlet distribution $\textrm{Dir}((\delta + n_{1i, l'})_{i = 1}^{l'})\times\textrm{Dir}((\delta + n_{2i, l'})_{i = 1}^{l'})$. Thus, the acceptance ratio is \normalsize \begin{equation} \begin{split} &Q(\Theta\rightarrow\Theta') = \min\{1, \frac{\psi(\Theta')g(\Theta|\Theta')}{\psi(\Theta)g(\Theta'|\Theta)}\} = \\ &\min\{1, \frac{\exp(\sigma l)\prod_{i = 1}^l|r_{i, l}|^{n_{1i, l} + n_{2i, l}}p(A_l, l|A_{l'}, l')\beta((\delta + n_{1i, l'})_{i = 1}^{l'})\beta((\delta + n_{2i, l'})_{i = 1}^{l'})}{\exp(\sigma l')\prod_{i = 1}^{l'}|r_{i, l'}|^{n_{1i, l'} + n_{2i, l'}}p(A_{l'}, l'|A_l, l)\beta((\delta + n_{1i, l})_{i = 1}^l)\beta((\delta + n_{2i, l})_{i = 1}^l)}\} \end{split} \label{acc_ratio} \end{equation} \normalsize where $\psi(\Theta|\mathcal{X}, \mathcal{Y})$ is defined in \eqref{post} and $\mathcal{X}, \mathcal{Y}$ are omitted to simplify notations. According to \eqref{acc_ratio}, $\mathbf{m}_{1, l}, \mathbf{m}_{1, l'}$ and $\mathbf{m}_{2, l}, \mathbf{m}_{2, l'}$ are canceled, $Q(\Theta\rightarrow\Theta')$ depends on $(A_l, l)$ and $(A_{l'}, l')$ solely and avoids to searching the vast space of probability masses. Moreover, if $p(l'|l)$ is chosen to be symmetric, $Q(\Theta\rightarrow\Theta')$ can be further simplified. In higher dimensions, sampling according to $p(A_{l'}|A_l, l')$ requires to count the number of points of $\mathcal{X}$, $\mathcal{Y}$ in each sub-rectangle of $A_{l'}$, which is expensive (with complexity $O((n_1 + n_2)d)$). If we are willing to run longer chains with cheaper cost per iteration, another heuristic choice for transition $p(A_{l'}, l'|A_l, l)$ is to increase or shrink $A_l$ by one uniformly: $p(A_{l + 1}, l + 1|A_l, l) = p(l + 1|l) / (l\cdot d)$ since there are $l\cdot d$ possible decisions (cutting locations), $p(A_{l - 1}, l - 1|A_l, l) = p(l - 1|l)$ and $p(l + 1|l) + p(l - 1|l) = 1$ when $l > 1$; $p(A_{l + 1}, l + 1|A_l, l) = 1 / (l\cdot d)$ when $l = 1$. \begin{figure} \centering \includegraphics[width=.6\textwidth]{joint_par_sim_1.pdf} \caption{\textbf{Sanity Tests.} First row: 1,000 samples drawn from $p_{11}$ and $p_{21}$ and the learned partition; second row: 1,000 samples drawn from $p_{21}$ and $p_{22}$ and the learned partition.} \label{joint_sim} \end{figure} \section{Numeric Simulations} \subsection{Sanity Tests for co-BPM} We use two ``sanity tests'' to demonstrate that our model is sensitive to differences among samples, the densities used in experiments are two dimensional for ease of visualization. In the first group $p_{11}$ and $p_{12}$, we force the mixture of $p_{11}$ and $p_{12}$ is uniformly distributed, i.e., $\frac{1}{2}(p_{11} + p_{12}) = \mathbf{1}\{x\in [0, 1]^2\}$; we generate 1,000 points each, thus the combined sample is equivalent to 2,000 points from uniform distribution; as shown in the first row of Figure \ref{joint_sim}, their difference is revealed by the partition well. The second group $p_{21}$ and $p_{22}$ are chosen such that they are both supported on the binary partition and $p_{22}$ is defined on a finer partition than $p_{21}$; a good model should be able to produce the underlying partition, i.e., find a density that is finer than both densities; the second row of Figure \ref{joint_sim} shows this result. \normalsize \[p_{11}(x) = \frac{9}{5}\mathbf{1}\{x\in [0, 1]^2\} - \frac{4}{5}\beta_{2, 2}(x_1)\beta_{2, 2}(x_2)\] \[p_{12}(x) = \frac{1}{5}\mathbf{1}\{x\in [0, 1]^2\} + \frac{4}{5}\beta_{2, 2}(x_1)\beta_{2, 2}(x_2)\] \[p_{21}(x) = \frac{2}{3}\mathbf{1}\{x\in r_1\} + \frac{4}{3}\mathbf{1}\{x\in r_2\} + \mathbf{1}\{x\in r_3\cup r_4\cup r_5\cup r_6\}\] \[p_{22}(x) = \frac{2}{3}\mathbf{1}\{x\in r_1\} + \frac{4}{3}\mathbf{1}\{x\in r_2\} + \frac{1}{2}\mathbf{1}\{x\in r_3\} + \mathbf{1}\{x\in r_4\} + \frac{4}{3}\mathbf{1}\{x\in r_5\} + \frac{8}{3}\mathbf{1}\{x\in r_6\}\] \normalsize \subsection{Simple Examples in 1 and 3 Dimensions} We demonstrate our methods by 2 simulations with dimension $d = 1, 3$ and size $n_1 = n_2 = 50, 250, 1250$. The parameters for co-BPM are $\delta = 1/2$ (which is the Jeffrey's non-informative prior), $\sigma = d + 1$, $p(l + 1|l) = p(l - 1|l) = 1/2$ for $l > 1$. The number of replicas for the box-plot is 3,000 and the burn-in number is 5,000. The discrepancies we consider are Total Variation, Hellinger Distance and KL divergence and $\alpha-$divergence with $\alpha = 2$. The true values of the discrepancies are obtained by Monte Carlo method with $10^8$ samples.\\ \textbf{1-dimensional examples.} The densities are defined as below and the results are summarized in Figure \ref{one_dim}. \normalsize\[p_1(x) = \beta_{6, 5}(x), p_2(x) = \beta_{5, 6}(x)\]\normalsize where $\beta_{a, b}(x)$ is the Beta distribution with shape parameters $a, b$. \begin{figure} \centering \includegraphics[width = .8\textwidth]{one_dim.pdf} \caption{\textbf{1-dimensional simulation.} We draw 8,000 samples and generate box-plot after discarding the first 5,000 burn-in sample points. Theoretical discrepancies (red lines): 0.2461, 0.2207, 0.2000, 0.4056.} \label{one_dim} \end{figure}\\ \textbf{3-dimensional examples.} The densities are defined as below and the results are summarized in Figure \ref{three_dim}. \begin{equation} \begin{gathered}p_1(x, y, z) = \frac{2}{5}\beta_{1, 2}(x)\beta_{2, 3}(y)\beta_{3, 4}(z) + \frac{3}{5}\beta_{4, 3}(x)\beta_{3, 2}(y)\beta_{2, 1}(z) \\ p_2(x, y, z) = \frac{2}{5}\beta_{1, 3}(x)\beta_{3, 5}(y)\beta_{5, 7}(z) + \frac{3}{5}\beta_{7, 5}(x)\beta_{5, 3}(y)\beta_{3, 1}(z) \end{gathered} \label{3_example} \end{equation} \begin{figure} \centering \includegraphics[width = .8\textwidth]{three_dim.pdf} \caption{\textbf{3-dimensional simulation.} First row, box-plot with the same configuration as 1-dimensional case; theoretical discrepancies (red lines): 0.2301, 0.2129, 0.2133, 0.6769. Second row, we assess the behavior of depth $l$ after discarding first 5,000 burn-in samples, the increase of number of sub-rectangles as sample size increases indicates that co-BPM refines the partitions when more information becomes available, which is analogous to the multi-resolution property discussed in \citep{Wong2010}.} \label{three_dim} \end{figure}\\ It is observed from Figure \ref{one_dim} and \ref{three_dim} that co-BPM estimates the divergences reasonably well. Moreover, the estimation errors and variances are decreased when the sample size increases, so is the number of outliers in box-plots. According the histogram of depth $l$ in Figure \ref{three_dim}, another interesting observation is that the number of sub-rectangles increases in tandem with the sample size, which indicates that co-BPM refines the partitions to reveal more structure of the samples when more information becomes available, this multi-resolution property of binary partition is also discussed in Optional P\'{o}lya Tree \citep{Wong2010}. The vanishing boundaries of beta distribution cause the large range and variance of KL and $\alpha-$divergence where division is involved, as demonstrated by the previous examples. Instead of using Jeffrey's noninformative prior, we choose a stronger prior with larger $\delta$ as a tradeoff between bias with variance. According to law of total variance, $\mathrm{Var}(m_{1i}) = E[\mathrm{Var}(m_{1i}|\mathcal{X})] + \mathrm{Var}[E(m_{1i}|\mathcal{X})]$. Since $m_{1i}$ is generated from Dirichlet distribution, $\mathrm{Var}(m_{1i}|\mathcal{X}) = \frac{(n_{1i} + \delta)(n_1 + \delta l - n_{1i} - \delta)}{(n_1 + \delta l)^2(n_1 + \delta l + 1)} = O(\delta^{-1})$ and $\mathrm{Var}[E(m_{1i}|\mathcal{X})] = \mathrm{Var}(\frac{n_{1i} + \delta}{n_1 + \delta l}) = \frac{\mathrm{Var}(n_{1i})}{(n_1 + \delta l)^2}$; thus, asymptotically, its variance is reduced with a larger $\delta$, as shown in Table \ref{four_dim}. \begin{table} \center \begin{tabular}{cc} \putindeepbox[10pt]{\includegraphics[width=0.25\textwidth]{bias_var.pdf}} & \putindeepbox[10pt]{\scalebox{1.3}{\begin{tabular}{@{}r|rrrr} \toprule & \multicolumn{4}{c}{Stronger Prior ($\delta$)}\\%&\multicolumn{4}{c}{Uniform Samples ($\pi$)}\\ \hline & $0.5$ & $0.6$ & $0.7$ & $0.8$\\ \midrule \hline Medians & 2.3057 & 1.7771 & 1.5960 & 1.5267\\ Std & 1.2690 & 0.6643 & 0.2223 & 0.2494\\ \bottomrule \end{tabular}}} \end{tabular} \caption{\textbf{Illustration of bias and variance tradeoff.} $\mathcal{X}\sim\mathcal{N}(\mu_1\mathbf{1}, \sigma_1^2\mathbf{I})$, $\mathcal{Y}\sim\mathcal{N}(\mu_2\mathbf{1}, \sigma_2^2\mathbf{I})$ and are truncated in $[0, 1]^4$, where $|\mathcal{X}| = |\mathcal{Y}| = 500$; $\mu_1 = 1/3, \mu_2 = 1/2$ and $\sigma_1 = 1/5, \sigma_2 = 1/5$; $\mathbf{1}$ is 4-dim unit vector and $\mathbf{I}$ is 4-dim identity matrix. As pseudo-count $\sigma$ increases, the variance decreases and bias increases. True value: 2.2196. The table on the right lists the medians and standard errors.} \label{four_dim} \end{table} \subsection{Comparison with Other Methods} Most of other methods focus on KL divergence estimation. \citep{Nguyen2010} derive a general lower bound for $f-$divergence via conjugate dual function; however, all of their theories and experiments are based on KL divergence estimation in 1, 2, 3 dimensions and their algorithms require that two samples have same size. In this section, we compare co-BPM to the methods in \citep{Nguyen2010} and \citep{Perez-Cruz2008} in KL divergence estimation. We briefly describe the methods in \citep{Nguyen2010} and \citep{Perez-Cruz2008}, interested readers may refer to the original papers for details. The core of \citep{Nguyen2010}'s algorithm is a convex program that has the number of parameters equal to the sample size. As pointed out by the authors, the performance of the algorithm depends on a regularization parameter $\lambda$; here we choose $\lambda$ according to their suggestions. There are two slightly different versions of KL estimator proposed, which are denoted by NWJ-M1 and NWJ-M2 in our experiments. On the other hand, the idea of \citep{Perez-Cruz2008}'s approach is relatively straightforward: the \emph{$k-$nearest neighbor density estimate} is computed for each sample point in $\mathcal{X}$, i.e., $\hat{p}_{1}(x)$ and $\hat{p}_{2}(x)$ for $x\in\mathcal{X}$, then KL divergence is estimated by $\frac{1}{|\mathcal{X}|}\sum_{x\in\mathcal{X}}\log\frac{\hat{p}_{1}(x)}{\hat{p}_{2}(x)}$ (denoted by PC$-k$). Some more sophisticated algorithms are proposed based on $k-$nearest neighbor, e.g., \citep{Poczos2012}. The tuning parameter $k$ is critical to guarantee their performance, we follow the choice of \citep{Perez-Cruz2008} by setting $k = 1$ and $k = 10$. One should also notice that the positivity of the estimates is not guaranteed. \begin{figure}[!ht] \center \includegraphics[width = 0.8\textwidth]{compdata.pdf} \caption{\textbf{Comparisons of NWJ, PC$-k$ and co-BPM.} All densities are truncated in $[0, 1]^3$ and each line is slightly shifted to avoid overlaps. Left: $p_1(x) \sim \mathbf{1}\{x\in [0, 1]^3\}$, $p_2(x) \sim \mathcal{N}(0, (1 / 3)^2\mathbf{I})$; Middle: $p_1(x) \sim \mathcal{N}(0, (1 / 2)^2\mathbf{I})$, $p_1(x) \sim \mathcal{N}(1, (1 / 2)^2\mathbf{I})$; Right: Beta mixture in \eqref{3_example}.} \label{comp} \end{figure} We report their performance in Figure \ref{comp}. For each of the three estimation problems described here, we experiment with 3 dimensional distributions and increasing sample sizes (the sample size, ranges from 50 to or more in logarithmic scale). Error bars are obtained by replicating each set-up 10 times. We see that co-BPM generally exhibits the best performance among the estimators considered. The estimates of NWJ-M1 and NWJ-M2 is somewhat less good and have larger variance; moreover, the convex program of NWJ-M1 or NWJ-M2 is computationally difficult because of the logarithm or entropy term and its scale increases rapidly with the sample size \citep{Grant2008}. However, PC$-k$ is more stable than NWJ and $k$ seems to strike a balance between bias and variance---the higher $k$ corresponds to lower variance but larger bias, e.g., its variance is smallest when $k = 10$ in Figure \ref{comp}; on the right plot, PC$-k$ produces negative estimates with small sample size since it can not guarantee positivity. \subsection{Examples with $\textrm{Dimensions} > 3$} There are few experiments in dimensions higher than 3 in previous work, even though most of the methods claim that there are no technical difficulties when dimensionality increases. The reason is because that their performance degrades fast as dimensionality increases. We demonstrate two additional examples with $d = 5, 10$ and assess the convergence rate empirically, the densities in our example are \normalsize\[p_1(x) = \frac{24}{25}\prod_{i = 1}^d\beta_{1, 5}(x_i) + \frac{1}{25}\mathbf{1}\{x\in [0, 1]^d\}, p_2(x) = \frac{49}{50}\prod_{i = 1}^d\beta_{5, 1}(x_i) + \frac{1}{50}\mathbf{1}\{x\in [0, 1]^d\}\]\normalsize The results are shown in Figure \ref{high_dim}, the sample size is plotted on the log scale and the error bars are standard deviation of 10 replicas. For KL divergence, the estimates of \citep{Nguyen2010} and \citep{Perez-Cruz2008} are also plotted for reference. There are several observations: (1) On the log scale, the estimates converge to true values almost linearly, and is comparable to the results of NWJ-M1; (2) NWJ-M1 is slightly better than NWJ-M2, which tends to underestimate the KL divergence. (3) PC$-k$ becomes inferior rapidly when the dimensionality goes up, especially in small sample size with large negative estimates. This is partially related to the sparseness in multi-dimensions which renders $k-$nearest neighbor density estimator ineffective \citep{Givens2012}. There is another noticeable phenomenon worth further investigation---the estimates goes down first then up as sample size increases. (4) Compared with previous examples, the estimates in dimensions $\leq 3$ converge relatively faster than those when dimensions are $5$ or $10$, this degradation of convergence corroborates the already cited work, which corresponds to the difficulty in multivariate density estimation. \begin{figure}[ht] \center \includegraphics[width = .8\textwidth]{higher_dim.pdf} \caption{ \textbf{Examples and comparison when dimensions $> 3$.} log(Sample size) vs. Divergences. Top: 5-dim example, the true values of Total Variation, Hellinger Distance, KL divergence and $\alpha-$divergence with $\alpha = 2$ are 0.9756, 0.9520, 7.6365, 8.9440 respectively. Bottom: 10-dim example, the true values are 0.9790, 0.9779, 11.3819, 13.8341.} \label{high_dim} \end{figure} \section{Conclusion and Discussion} A unified approach to estimating discrepancies of distributions is proposed from a Bayesian perspective. The joint prior provides a single-shot way to estimate the values and conduct statistical inferences; the simulations demonstrate its attractive empirical performance. In applications, this approach can be naturally extended to handle multiple-sample case, for example, given 3 samples $\mathcal{X}, \mathcal{Y}, \mathcal{Z}$ drawn from $p_1, p_2, p_3$ respectively and some distance measure $D(\cdot, \cdot, \cdot)$, we can estimate $D(p_1, p_2, p_3)$ according to $\mathcal{X}, \mathcal{Y}, \mathcal{Z}$ by binary-partitioning their sample space jointly when similar lower bound in Theorem \ref{thm1} exists. The core of our methodology is Theorem \ref{thm1}, which opens up the possibility to apply any domain partition based density estimators, a possible future direction is to use even more adaptive partitioning schemes. \bibliographystyle{apalike}
\section*{Abstract} A general circulation model of intermediate complexity with an idealized earthlike aquaplanet setup is used to study the impact of changes in the oceanic heat transport on the global atmospheric circulation. Focus is put on the Lorenz energy cycle and the atmospheric mean meridional circulation. The latter is analysed by means of the Kuo-Eliassen equation. The atmospheric heat transport compensates the imposed oceanic heat transport changes to a large extent in conjunction with significant modification of the general circulation. Up to a maximum about 3PW, an increase of the oceanic heat transport leads to an increase of the global mean near surface temperature and a decrease of its equator-to-pole gradient. For larger transports, the gradient is reduced further but the global mean remains approximately constant. This is linked to a cooling and a reversal of the temperature gradient in the tropics. A larger oceanic heat transport leads to a reduction of all reservoirs and conversions of the Lorenz energy cycle but of different relative magnitude for the individual components. The available potential energy of the zonal mean flow and its conversion to eddy available potential energy are affected most. Both Hadley and Ferrel cell show a decline for increasing oceanic heat transport with the Hadley cell being more sensitive. Both cells exhibit a poleward shift of their maxima, and the Hadley cell broadens for larger oceanic transports. The partitioning by means of the Kuo-Eliassen equation reveals that zonal mean diabatic heating and friction are the most important sources for changes of the Hadley cell while the behaviour of the Ferrell cell is mostly controlled by friction. \newpage \section{Introduction} Astronomical factors and differences of local albedo cause a difference of net incoming shortwave radiation between low and high latitudes leading to differential heating and a surplus of energy in the tropics. Considering global and long-term averages, at steady-state the same amount of supplied energy is emitted to space again, and the incoming shortwave radiation is balanced by the outgoing longwave radiation (Peixoto and Oort 1992, Lucarini and Ragone, 2011). Atmospheric and oceanic transports, fuelled by instabilities due to the presence of temperature differences related to inhomogeneous heating, tend to reduce such temperature differences, thus acting as a powerful negative feedback that stabilizes the climate system. Stone (1978) argued that the magnitude of the total meridional heat transport, i.e. the sum of the oceanic and the atmospheric contributions, is insensitive to the structure and the specific dynamical properties of the atmosphere-ocean system. That is, changes of the oceanic heat transport (OHT) will be compensated by the atmospheric flow and vice versa. In particular, he suggested that the peak of the heat transport is constrained within a narrow range of latitudes regardless of the radiative forcing. If the climate system is at steady-state, the features of the meridional heat transport can be related to the solar constant, the radius of the Earth, the tilt of the Earth's axis and the hemispheric mean albedo. Stone argued that the insensitivity to the structure and the dynamics of the system is due to the correlation of thermal emissions to space, the albedo and the efficiency of the transport mechanisms of the atmosphere and the ocean. Enderton and Marshall (2008) discussed the limits of Stone's hypothesis by employing a series of coupled atmosphere-ocean-sea ice model experiments in which the oceanic circulation on an aqua-planet is constrained by different meridional barriers. The presence or absence of the barriers results in significant different climates, in particular in climates with and without polar ice caps. They concluded that Stone's result is a good guide for ice free climates. But, if polar ice caps are present the effect of the related meridional gradients in albedo on the absorption of solar radiation need to be taken into account. The atmospheric compensation implies a significant impact of changes in OHT on the atmospheric circulation as a whole which affects the zonally symmetric flow, the zonally asymmetric (eddy) flow and the interplay between both. Thus, changes in OHT have been commonly used to account for paleo-climatic changes (e.g. Rind and Chandler 1991, Sloan et al. 2001, Romanova et al. 2006). Moreover, OHT is an important factor for a potential anthropogenic climate change since significant modifications of it can expected. Unfortunately, large uncertainties exist of changes in the oceanic circulation simulated in climate change scenarios (IPCC 2013). These result from, amongst others, the uncertainties in fresh water forcing due to potential melting of inland ice sheets. To assess the role of the ocean for historical and potential future climates the impact of the OHT on the atmospheric circulation and the underlying mechanisms need to be investigated systematically. A way of studying the impact of changes in OHT on the atmospheric circulation is to utilize an atmospheric general circulation model coupled to a mixed-layer ocean. In such a model the OHT can be prescribed. Using a present-day setup including continents Winton (2000), Herweijer et al.~(2005) and Barreiro et al.~(2011) found that increasing OHT results in a warmer climate with less sea ice. A reduction of low level clouds and an increase of greenhouse trapping due to a moistening of the atmosphere appeared to be relevant mechanisms. In addition, a weakening of the Hadley cell with increased OHT was found by Herweijer et al.~and Barreiro et al. Utilizing an idealized aqua-planet setup Rose and Ferreira (2013) systematically assessed the impact of the OHT on the atmospheric global mean near surface temperature and its equator-to-pole gradient. For warm and ice-free climates they confirm a near-perfect atmospheric compensation of the imposed changes in OHT. Like in the above studies including continents, they found they found an increase in global mean temperature for increasing OHT, accompanied by a decrease in the equator-to-pole temperature gradient. Tropical SSTs showed to be less affected than higher altitudes. The detailed meridional structure of the oceanic heat transport turned out to be less important. Changes in deep moist convection in the mid-latitudes together with an enhanced water vapor greenhouse appear to be the major drivers. Koll and Abbot (2013) confirmed the low sensitivity of tropical SSTs to OHT changes. In their aqua-planet experiments larger OHT leads to a weakening of the Hadley cell which reduces cloud cover and surface winds, and, thus, counteracts surface cooling by increased OHT. In the present study we extend and supplement the above studies. Based on the experimental setup of Rose and Ferreira our analyses focus on the impact of OHT changes on the atmospheric dynamics. The integrated effect on the atmospheric energetics is assessed by means of the Lorenz energy cycle. Changes in the atmospheric mean meridional circulation are analysed employing the Kuo-Eliassen equation. In Section 2 we describe the model and the experimental design. Section 3 introduces our diagnostics. The results of the analyses are presented in Section 4. A summary and discussion concludes the paper (Section 5). \section{Model and experimental setup} The Planet Simulator (PlaSim, \textit{http://www.mi.uni-hamburg.de/Planet-Simul.216.0.html}) is an open source general circulation model (GCM) of intermediate complexity developed at the University of Hamburg. For the atmosphere, the dynamical core is the Portable University Model of the Atmosphere (PUMA) based on the primitive equation multi-level spectral model of Hoskins and Simmons (1975) and James and Gray (1986). Radiation is parameterized by differentiating between shortwave and longwave radiation and between a clear or a cloudy atmosphere following the works of Lacis and Hansen (1974) for the short wave part, Sasamori (1968) for the long-wave, and Stephens (1978) and Stephens et al. (1984) for the radiative properties of clouds. Cloud fraction is computed according to Slingo and Slingo (1991). The representation of boundary-layer fluxes and of vertical and horizontal diffusion follows Louis (1979), Louis et al.(1982), Roeckner et al. (1992) and Laursen and Eliasen (1989). The cumulus convection scheme is based on Kuo (1965, 1974). The ocean is represented by a thermodynamic mixed-layer (slab ocean) model including a 0-dimensional thermodynamic sea-ice. Following Rose and Ferreira (2013) we use an earthlike aqua-planet setup with zonally symmetric forcing utilizing present day conditions for the solar constant (1365Wm$^{-2}$) and the CO$_2$-concentration (360ppm). The solar insolation comprises an annual cycle but eccentricity is set to zero. Thus, on annual average the forcing is hemispherically symmetric as well. The mixed layer depth is set to $60\,m$. A temporally constant flux into the ocean (q-flux) is used to prescribe the oceanic heat transport (OHT) according to the analytic equation given by Rose and Ferreira (2013): \begin{equation} \mbox{OHT}=OHT_0 \cdot \sin(\phi)\cos(\phi)^{2N} \end{equation} Where $\phi$ denotes the latitude. $N$ is a positive integer which determines the latitude of the maximum of the transport and the shape of its meridional profile, and $OHT_0$ is a constant defining the magnitude. Rose and Ferreira made sensitivity experiments by varying $N$ (ranging from 1 to 8) and by varying the peak transport (ranging from 0PW to 4PW) which is controlled by $OHT_0$. For our study we follow Rose and Ferreira but fix the location of the peak by setting $N=2$ (which corresponds to maximum transport at 27$^{\circ}$). We perform nine sensitivity simulations with respect to the magnitude of the transport by changing $OHT_0$ to obtain peak transports OHT$_{max}$ from 0PW to 4PW (with 0.5PW increment). OHT$_{max}$=0PW (i.e. no OHT) serves as the control simulation. The OHT for OHT$_{max}$=0,1,2,3 and 4PW is displayed in Figure \ref{fig:OHT}. All simulations are run for at least 100 years (360 days per year). The last 30 years are subject to the analyses. A horizontal resolution of $T31$ ($96\,\times\,48$ grid points) with five $\sigma$-levels in the vertical is used. The timestep is $\Delta t=23\,min$. \section{Diagnostics} The analyses focus on the Lorenz energy cycle and the mean meridional circulation. The latter will be studied by means of the Kuo-Eliassen equation. Both diagnostics are briefly introduced in the following. \subsection{The Lorenz energy cycle} \label{sec:Lorenz} The atmospheric energy cycle proposed by Lorenz (1955) is one of the most important concepts to understand the global atmospheric circulation by means of energy conservation and by considering the integrated effects of physical mechanisms involved, that is, e.g., the generation of available potential energy by external forcing, the dissipation of kinetic energy and the energy conversions by baroclinic and barotropic processes. At the same time it gives information about the relative importance of the zonal mean circulation, the eddies and the interaction between both. Refering to the reservoirs of zonal available potential energy, eddy available potential energy, zonal kinetic energy and eddy kinetic energy as $P_M$, $P_E$, $K_M$ and $K_E$, respectively, the Lorenz energy cycle (i.e. the budget equations) may be written as \begin{eqnarray*} \frac{dP_M}{dt} &=& [S_P]-C(P_M,P_E)-C(P_M,K_M) \\ \frac{dP_E}{dt} &=& S_P^*+C(P_M,P_E)-C(P_E,K_E) \\ \frac{dK_E}{dt} &=& S_E^*+C(P_E,K_E)-C(K_E,K_M) \\ \frac{dK_M}{dt} &=& [S_E]+C(K_E,K_M)+C(P_M,K_M) \\ \end{eqnarray*} where $[S_P]$, $S_P^*$, $[S_E]$ and $S_E^*$ are external sources/sinks of the respective quantities and $C(A,B)$ denotes the conversion from $A$ to $B$. To compute the individual contributions we follow the work of Ulbrich and Speth (1991). In pressure coordinates, the reservoirs are given by \begin{eqnarray*} P_M &=& \langle \frac{\gamma}{2}([T]-\{T\})^2 \rangle \\ P_E &=& \langle \frac{\gamma}{2}[T^{*2}]\rangle \\ K_M &=& \langle \frac{1}{2}([u]^2+[v]^2) \rangle \\ K_E &=& \langle \frac{1}{2}([u^{*2}]+[v^{*2}]) \rangle \\ \end{eqnarray*} and the conversion terms are \begin{eqnarray*} C(P_M,P_E) &=& -\left\langle \gamma \left( [v^*T^*]\frac{\partial[T]}{r\partial \phi} +[\omega^* T^*]\left( \frac{\partial ([T]-\{T\})}{\partial p}- \frac{R}{p \cdot cp}([T]-\{T\})\right) \right) \right\rangle \\ C(P_M,K_M) &=& - \left\langle \frac{R}{p}([\omega]-\{\omega\}) ([T_v]- \{T_v\}) \right\rangle \\ C(P_E,K_E) &=& - \left\langle \frac{R}{p}[\omega^* T^*_v] \right\rangle \\ C(K_M,K_E) &=& \left\langle \left( [u^*v^*]\frac{\partial[u]}{r \partial \phi} +[u^*v^*][u]\frac{tg \phi}{r} +[v^*v^*]\frac{\partial[v]}{r \partial \phi} \right. \right. \\ & & \left. \left. -[u^*u^*][v]\frac{tg \phi}{r} +[\omega^*u^*]\frac{\partial[u]}{\partial p} +[\omega^*v^*]\frac{\partial[v]}{\partial p} \right) \right\rangle \\ \end{eqnarray*} with {\allowdisplaybreaks \begin{eqnarray*} [x] &=& \mbox{zonal mean} \\ x^* &=& \mbox{deviation from zonal mean} \\ \{x\} &=& \mbox{global horizontal mean} \\ <x> &=& \frac{1}{g\cdot A} \int_A\int_p x dp dA \\ A &=& \mbox{horizontal Area} \\ cp &=& \mbox{specific heat at const.~pressure} \\ g &=& \mbox{gravity} \\ p &=& \mbox{pressure} \\ r &=& \mbox{radius of the Earth} \\ R &=& \mbox{gas constant} \\ T &=& \mbox{temperature} \\ T_v &=& \mbox{virtual temperature} \\ u &=& \mbox{zonal wind} \\ v &=& \mbox{meridional wind} \\ \omega &=& \mbox{vertical (p) velocity}\\ \phi &=& \mbox{latitude} \\ \gamma &=& \mbox{stability parameter} = -\frac{R}{p}\left(\frac{\partial [\overline{T}]}{\partial p} -\frac{R}{cp}\frac{[\overline{T}]}{p}\right)^{-1} \\ \end{eqnarray*} } The external sources/sinks are diagnosed from the respective residuals. We note that in Ulbrich and Speth this energetics was formulated for mixed space-time domain. In our case, however, the contributions by stationary eddies is zero because of the zonally symmetric forcing. \subsection{The mean meridional circulation} To analyse the mean meridional circulation we make use of the so called Kuo-Eliassen equation (Kuo 1956, Eliassen 1951). It is a diagnostic equation which relates the mean meridional circulation (i.e. Hadley, Ferrel and Polar cell) to different sources. Applying the quasi-geostrophic approximation and defining a streamfunction $\psi$ with \begin{eqnarray*} [v] & = & \frac{g}{2\pi r \cos \phi}\frac{\partial \psi}{\partial p} \\ {[}\omega] &=& -\frac{g}{2\pi r \cos \phi}\frac{\partial \psi}{r\partial \phi} \\ \end{eqnarray*} the Kuo-Eliassen equation may be derived as (see, e.g., Peixoto and Oort 1992 chapter 14.5.5) \begin{eqnarray*} \frac{f^2g}{2\pi r \cos \phi}\frac{\partial^2 \psi}{\partial p^2} -\frac{g}{2\pi r^3 \rho [\theta ]}\frac{\partial}{\partial \phi} \left(\frac{\partial [\theta ]}{\partial p}\frac{\partial \psi}{\partial \phi} \right) & = & \frac{1}{r\rho [T]}\frac{\partial}{\partial \phi}\frac{[Q]}{cp} \\ & - & f \frac{\partial[F]}{\partial p} \\ & - & \frac{1}{r^2\rho [\theta ]} \frac{\partial}{\partial \phi} \frac{\partial [v^*\theta^*]\cos\phi}{\cos\phi\partial\phi} \\ & + & \frac{f}{r\cos^2\phi}\frac{\partial^2[u^*v^*]\cos^2\phi}{\partial p \partial\phi} \\ \end{eqnarray*} Where, in addition to the symbols defined above, $f$ is the Coriolis parameter, $\rho$ density, $\theta$ potential temperature, $Q$ diabatic heating and $F$ the tendency of the zonal wind due to friction. We solve the Kuo-Eliassen equation for $\psi$ by applying an iterative method (Gauss-Seidel method) to its finite difference approximation. Thus, we are able to diagnose the contributions from the different sources to the mean meridional circulation, which are diabatic heating (1$^{st}$ term r.h.s), friction (2$^{nd}$), meridional eddy heat transport (3$^{rd}$) and eddy momentum transport (4$^{th}$). We note that though the equation, in the present form, involves the quasi-geostrophic approximation it has shown to be reasonably applicable even in the deep tropics (Kim and Lee 2001a,b). \section{Results} Before discussing the impact of changes in OHT on the Lorenz energy cycle and the mean meridional circulation we present the effect on the mean climate in terms of the total meridional heat transport and the atmospheric near surface (2m) temperature. First we note that similar to Rose and Ferreira our model exhibits multiple equilibria, a warm state and a snow ball Earth depending on the initial conditions as thoroughly discussed in Boschi et al.~(2013). In the present study we investigate the warm states only. However, in contrast to Rose and Ferreira sea ice at high latitudes is present in those warm state in all simulations. Despite the difference in sea ice extend (i.e. planetary albedo) the atmospheric heat transport compensates the changes in OHT to a great extend as can be seen from the total meridional heat transport diagnosed from the energy budged at the top of the atmosphere (Figure \ref{fig:THT}). Up to about OHT$_{max}$=2.5PW increasing OHT leads to an increase of the global mean ($T_{M}$) and a decrease of the equator-to-pole gradient ($\Delta T$) of the annual and zonal mean near surface air temperature (Figure \ref{fig:T2M+TMDT}). For this regime an approximately linear relationship between $T_{Mean}$ and $\Delta T$ can be found. For OHT$_{max}$>2.5PW, $T_{M}$ is almost insensitive to an OHT change while $\Delta T$ is further reduced when stronger OHTs are considered. Here, the equator-to-pole gradient is defined by the difference between the values at the lowest and highest latitude of the models grid. Inspecting the respective meridional profiles of the annual and zonal mean near surface temperatures we observe that high latitudes are more sensitive to the OHT changes than low latitudes. With increasing OHT the polar temperatures continuously increase except for OHT$_{max}$=4PW where slightly colder polar temperatures than for OHT$_{max}$=3.5PW are found. It appears that this is a consequence of the reduced atmospheric heat transport slightly over-compensating the increased but still small oceanic heat transport at these latitudes. In the tropics, an increase of the temperatures is only present until OHT$_{max}$=1.5PW. For larger OHT the equatorial temperatures decrease. In addition, increasing OHT leads to a flattening of the temperature profile in the tropics until, for OHT OHT$_{max}$=3.5 and 4PW, the temperature gradient in the tropics gets reversed and the maximum of the temperature shifts away from the equator to approx. $\pm 24\,^{\circ}$. Qualitatively, all findings are also true for winter and summer as can be seen in Figures \ref{fig:T2MJJA}, except that in summer the sensitivity to OHT changes is small in high latitudes which are covered by sea ice. In addition we note that the seasonality and its sensitivity to OHT changes are small for latitudes without sea ice due to the high thermal inertia of the mixed layer. \subsection{The Lorenz energy cycle} We compute the climatological average Lorenz energy cycle from 30 year daily data for the entire year and for June-August showing the respective terms for the winter (northern) and the summer (southern) hemisphere separately (Figure \ref{fig:LORENZ}). We note that by using the equations given in Section \ref{sec:Lorenz} the computed averaged values include contributions from the annual cycle. This, however, only affects the annual values of the reservoirs $P_M$ and $K_M$, and of the conversion $C(P_M,K_M)$. Using annual averaged values would diminish the annual $P_M$, $K_M$ and $C(P_M,K_M)$ by about 50\% each. The monotonic but nonlinear decrease of the global mean zonal averaged temperature gradient with increasing OHT which is present throughout the troposphere is reflected in the mean flow available potential energy $P_M$ (Figures \ref{fig:LORENZ}a,c,e). Compared to the OHT$_{max}$=0PW simulation, $P_M$ is reduced by about 71\% in the OHT$_{max}$=4PW run for yearly values, 73\% for the winter and 67\% for the summer hemisphere. A decrease with increasing OHT can also be detected for all other reservoirs. However, here the relative decrease for $P_E$ and $K_M$ is substantially smaller for the summer (approx.~48\% and 59\%, resp.) than for the winter hemisphere (approx.~77\% and 72\%). The yearly values reduced by approx.~69\% and 65\%. $K_E$ declines by about 54\% for the yearly data and for both hemispheres, thus showing by far the smallest decrease of all reservoirs on the winter hemisphere. The overall decline in the reservoirs with increasing OHT is also present for conversion terms which are mainly related to the activity of baroclinic synoptic eddies, i.e. $C(P_M,P_E)$, $C(P_E,K_E)$ and $C(K_E,K_M)$ (Figures \ref{fig:LORENZ}b,d,f). Here, the largest changes occur in winter where baroclincity (i.e. the temperature gradient) is reduced most. However, the sensitivity appears to decrease following the sequence of a baroclinic live cycle. While $C(P_M,P_E)$ show the largest sensitivity (approx.~65\%, 70\% and 47\% for annual, winter and summer, resp.), the changes in $C(K_E,K_M)$ are the smallest (approx.~53\%, 59\% and 34\%). The changes in $C(P_E,K_E)$ amount to approx.~57\%, 61\% and 44\%,respectively. The convergences of the conversions with increasing OHT indicate that zonally asymmetric diabatic heating and friction become less important for the Lorenz energy cycle. We note that in the present simulations the zonally asymmetric heating extracts $P_E$, i.e.~it acts to homogenize the zonal temperature profiles. For OHT$_{max}$>4PW there is almost no contribution of zonally asymmetric diabatic heating similar to the summer hemisphere in all simulations. $C(P_M,K_M)$ exhibits the smallest sensitivity. For the summer hemisphere, a decrease with increasing OHT can be found indicating that the thermally indirect zonal mean circulation (Ferrel cell) becomes more important compared to the thermally direct one (Hadley Cell). For the winter hemisphere, the thermally direct mean circulation dominates in all simulations but with slightly decreasing magnitude. On annual basis little changes in $C(P_M,K_M)$ are found. In summary, an increase in OHT leads to a decrease in all components of the Lorenz energy cycle. This suggests that the atmosphere gets less active and less efficient. We note that the strength of the Lorenz energy cycle can be linked to a Carnot efficiency of the climate system (Lucarini 2009). Indeed, this efficiency declines with increasing OHT. A study on this efficiency, and on other global thermodynamic properties, will be presented in a companion paper. It turns out that changes in the intensity of the water vapor transport are most relevant, similar to Lucarini et al.~(2010). \subsection{The mean meridional circulation} For discussing the changes in the mean meridional circulation we focus on the climatological annual average of the zonal mean mass stream function ($\psi$). However, if not stated otherwise, the results are qualitatively similar for all seasons. Figure \ref{fig:MMC} shows northern hemisphere $\psi$ for OHT$_{max}$=0, 2, 3 and 4PW. For OHT$_{max}$=0PW, a Hadley cell and a Ferrel cell are well established with values of about 8$\cdot$10$^{10}$kg s$^{-1}$ and -3$\cdot$10$^{10}$kg s$^{-1}$, respectively. The maximum magnitudes are located at about 10$^{\circ}$N for the Hadley cell and 50$^{\circ}$N for the Ferrel cell, and at about 700hPa for both cells. The Hadley cell extends to about 33$^{\circ}$N. A polar cell is absent in the annual mean but emerges weakly in the summer months. With increasing OHT the strength of both cells decreases (Figure \ref{fig:HFC}). However, while the strength of the Hadley cell is virtually linear and amounts to about 85\%, the Ferrel cell strength decreases by about 50\% only with stronger decrease for smaller OHT$_{max}$. In addition to the changes in strength, the Ferrel cell shifts poleward. For OHT$_{max}$>2PW a poleward shift can also be observed for the Hadley cells maximum together with a broadening of this cell, i.e.~a poleward shift of its edge. For OHT$_{max}$=4PW an additional thermally indirect cell can be observed close to the equator. This is related to an almost vanished Hadley cell in summer together with a winter hemisphere Hadley cell which has its maximum on the summer hemisphere. The reconstructions by means of the Kuo-Eliassen equation are in good agreement with the actual $\psi$ for all simulations even if the maximum magnitudes are somewhat overestimated. This gives us confidence to the applied methodology. Figure \ref{fig:MMCR} shows the sources and the reconstruction for OHT$_{max}$=0PW. The individual sources indicate that the largest contributions to $\psi$ stem from diabatic heating and from friction. The heating controls the Hadley cell together with a significant contribution by friction. For the Ferrel cell friction is most important. To a much smaller extent the eddy transports of heat and of momentum add to the Ferrel circulation with a larger contribution by the heat transport. For both Hadley and Ferrel cell the maximum contribution by friction is located at lower levels than for all other sources. For the Hadley cell both major sources, heating and friction, decrease linearly with increasing OHT (Figure \ref{fig:MMCF}). As the decrease is stronger for heating, friction becomes the major source for OHT$_{max}$>3PW. The decrease of the Ferrel cell with increasing OHT is linked to a decrease of the friction, i.h. a decrease of the near surface zonal mean zonal wind. The contributions by the heat and momentum transports decrease less and remain constant for OHT$_{max}$>2PW. Similar to the changes of magnitude, the shifting of the cells and the broadening of the Hadley cell can be explaind by respective changes in the mean sources. \section{Summary and discussion} We have studied the impact of the oceanic heat transport (OHT) on the atmospheric circulation focusing on the Lorenz energy cycle and the mean meridional circulation. Utilizing a general circulation model of intermediate complexity (PlaSim) which includes an oceanic mixed layer we have adopted an experimental design from Rose and Ferreira (2013). Here, an imposed oceanic heat transport of simple analytic form and with varying strength allows for systematic analyses. The meridional circulation has been studied by means of the Kuo-Eliassen equations which separates contributions by different sources: diabatic heating, friction and the eddy transport of heat and momentum. We found a compensation of the changes in oceanic heat transport by the atmosphere confirming Stones (1978) conclusions. The presence of sea ice may explain the deviations from a perfect compensation as discussed in Enderton and Marshall (2008). In agreement with Rose and Ferreira we have found an increase of the global mean near surface temperature and a decrease of the equator-to-pole temperature gradient with increasing OHT for OHT$_{max}$<3PW. For larger OHT the temperature gradient is still decreasing but the global average remains constant. For the tropics, there is a significant decrease of both temperature and its gradient for OHT$_{max}$>2PW with a reversal of the gradient for OHT$_{max}$>3PW. For smaller OHT we observed a slight warming and a reduction of the gradient with increasing OHT which is consistent with results from Koll and Abbot (2013). However, in their aqua-planet the tropical temperature increases for all imposed (positive) OHTs (up to 3PW). A tropical cooling for imposed oceanic heat transports somewhat larger than present-day values has also been found by Barreiro et al.~(2011) in a more complex coupled atmosphere-slab ocean model with present-day land-sea distribution suggesting that present-day climate is close to a state were the warming effect of OHT is maximized. Barreiro et al.~related the tropical cooling to a strong cloud-SST feedback and showed that the results are sensitive to the particular parameterizations. With respect to the Lorenz energy cycle, the decline of the meridional temperature gradient with increasing OHT is directly linked to a reduction of the available potential energy of the zonal mean flow ($P_M$). In addition, the reduced gradient indicates less baroclinicity as well as a weaker zonal mean zonal wind (jet) due to the the thermal wind relation. Finally, less baroclinicity hint to a reduced eddy activity. Thus, it is consistent that all reservoirs and conversions of the Lorenz energy cycle decrease with increasing OHT. However, the sensitivities differ. $P_M$ and the conversion from $P_M$ to $P_E$ exhibit the largest changes. Eddy kinetic energy, the barotropic conversion from eddy kinetic energy to zonal mean kinetic energy, and the conversion from zonal mean kinetic energy to $P_M$ are least affected. Confirming the results of previous studies (Herweijer et al.~(2005), Barreiro et al., Koll and Abbot) we have found a decrease of the Hadley cell for increasing OHT. In addition, the Hadley cell broadens and its maximum shifts polward when OHT obtains large values (OHT$_{max}$>2.5PW). Separating individual sources by applying the Kuo-Eliassen equation showed that the characteristics of the Hadley cell can be explained by the mean meridional circulations related to the diabatic heating and, to a smaller extend, to the friction. In our simulations, the meridional circulation induced by friction also controls the behavior of the Ferrel cell. Eddy transports of heat and momentum appear to be less important. This is different from results by Kim and Lee (2001b) where the mean meridional circulation related to eddy fluxes account for about 50\% of the Ferrel cells strength. The coarse vertical resolution adopted from Rose and Ferreira may be responsible for a reduced eddy activity. Overall our study demonstrate the large impact of the oceanic heat transport on the atmospheric circulation which effects the zonally symmetric flow, the zonally asymmetric flow and the interaction between both. By reducing the meridional temperature gradient an increased oceanic heat transport weakens the Lorenz energy cycle and slows down and shifts the Hadley and the Ferrel cell. The reduction of the meridional gradient of the surface temperature is one of the major features of global warming. Lu et al. (2007) showed a consistent weakening and poleward expansion of the Hadley cell in IPCC AR4 simulations. Hence, changes in the oceanic heat transport may significantly modify the response of the atmospheric circulation to greenhouse warming. A weakening of the oceanic meridional overturning circulation as predicted by the majority of coupled ocean-atmosphere general circulation models (though with large uncertainties; IPCC 2013) would therefor act as a negative feedback mechanism. This negative feedback might become even more important when strong melting of inland ice sheets due to global warming is taken into account. The associated input of large amounts of fresh water has a huge potential to slow down the oceanic circulation. Finally, all changes discussed above can also be seen as a phenomenological fingerprint of modified global thermodynamic properties of the atmospheric flow like entropy production and an Carnot efficiency (Lucarini 2009). Analysing these properties will provide a more general perspective in the context of geophysical fluid dynamics. Results from such an investigation will be presented in a companion work (Schr\"oder et. al. 2014). \section*{References} Barreiro, M., Cherchi, A., and Masina, S.: Climate sensitivity to changes in ocean heat transport, J.~Climate, 24, 5015-5030,2011. Boschi, R., Lucarini, V., and Pascale, S.: Bistability of the climate around the habitable zone: a thermodynamic investigation, Icarus, 226, 1724-1742, 2012. Eliassen, A.: Slow frictionally controlled meridional circulation in a circular vortex, Astro.~Norv., 5, 19-60, 1951. Enderton, D.~and Marshall, J.: Explorations of atmosphere-ocean-ice climates on an aquaplanet and their meridional energy transports, J.~Atmos.~Sci., 66, 1593-1610, 2008. Herweijer, C., Seager, R., Winton, M., and Clement, A.: Why ocean heat transport warms the global mean climate, Tellus, 57A, 662-675, 2005. Holton, J. R., 1992: An Introduction to Dynamic Meteorology, International Geophysics Series, 48(3), Academic Press, 1992. Hoskins, B.J.~and Simmons, A.J.: A multi-layer spectral method and the semi-implicit method, Quart.~J.~Roy.~Meteorol.~Soc., 101, 637-655, 1975. IPCC: Climate Change 2013: The Physical Science Basis. Contribution of Working Group I to the Fifth Assessment Report of the Intergovernmental Panel on Climate Change [Stocker, T.F., D.~Qin, G.-K.~Plattner, M.~Tignor, S.K.~Allen, J.~Boschung, A.~Nauels, Y.~Xia, V.~Bex and P.M.~Midgley (eds.)]. Cambridge University Press, Cambridge, United Kingdom and New York, NY, USA, 1535pp, 2013. James I.N.~and Gray, L.J.: Concerning the effect of surface drag on the circulation of a planetary atmosphere, Quart.~J.~Roy.~Meteorol.~Soc., 112, 1231-1250, 1986. Kim, H.-K.~and Lee, S.: Hadley cell dynamics in a primitive equation model. Part I: Axisymmetric flow, J.~Atmos.~Sci., 58, 2845-2858, 2001a. Kim, H.-K.~and Lee, S.: Hadley cell dynamics in a primitive equation model. Part I: Nonaxisymmetric flow, J.~Atmos.~Sci., 58, 2859-2871, 2001b. Kuo, H.-L.: Forced and free meridional circulations in the atmosphere, J.~Meteor., 13, 561-568, 1956. Kuo, H.-L.: On formation and intensification of tropical cyclones through latent heat release by cumulus convection, J.~Atmos.~Sci., 22, 40-63, 1965. Kuo, H.-L.: Further studies of the parameterization of the influence of cumulus convection on large-scale flow, J.~Atmos.~Sci., 31, 1232-1240, 1974. Koll, D.D.B.~and Abbot, D.S.: Why tropical sea surface temperature is insensitive to ocean heat transport changes, J.~Climate, 26, 6742-6749, 2013. Lacis, A.A.~and Hansen, J.E.: A parameterization for the absorption of solar radiation on the Earth's atmosphere, J.~Atmos.~Sci., 31, 118-133, 1974. Laurson, L.~and Eliasen, E.: On the effects of the damping mechanisms in an atmospheric general circulation model, Tellus, 41A, 385-400, 1989. Lorenz, E.N.: Available potential energy and the maintenance of the general circulation, Tellus, 7, 157-167, 1955. Louis, J.F.: A parametric model of vertical eddy fluxes in the atmosphere, Boundary Layer Meteorology, 17, 187-202, 1979. Louis, J.F., Tiedtke, M., and Geleyn, M.: A short history of the PBL parameterisation at ECMWF, Proceedings, ECMWF workshop on planetary boundary layer parameterization, Reading, 25-27 Nov. 81, 59-80, 1982. Lu, J., Vecchi, G.A., and Reichler, T.: Expansion of the Hadley cell under global warming, Geophys.~Res.~Lett., 34, L06805, 2007. Lucarini, V.: Thermodynamic efficiency and entropy production in the climate system, Phys.~Rev.~E, 80, 021118, 2009. Lucarini, V., Fraedrich, K., and Lunkeit, F.: Thermodynamics of climate change: generalized sensitivities, Atmos.~Chem.~Phys., 10, 9729-9737, 2010. Lucarini, V.~and Ragone, F.: Energetics of climate models: Net energy balance and meridional enthalpy transport, Rev.~Geophys., 49, RG1001, doi:10.1029/2009RG000323, 2011. Peixoto, J.P.~and Oort, A.H.: Physics of Climate, American Institute of Physics, 1992. Rind, D.~and Chandler, M.: Increased ocean heat transports and warmer climate, J.~Geophys.~Res., 96, 7437-7461, 1991. Roeckner, E., Arpe, K., Bengtsson, L., Brinkop, S., D\"umenil, L., Esch, M., Kirk, E., Lunkeit, F., Ponater, M., Rockel, B., Sausen, R., Schlese, U., Schubert S., and Windelband, M.: Simulation of the present-day climate with the ECHAM-3 model: Impact of model physics and resolution, Max-Planck Institut f\"ur Meteorologie, Report No. 93, 171pp, 1992. Romanova, V., Lohmann, G., Grosfeld, K., and Butzin, M: The relative roles of oceanic heat transport and orography on glacial climate, Quat.~Sci.~Rev., 25, 832-845, 2006. Rose, B.~and Ferreira, D.: Ocean heat transport and water vapor greenhouse in a warm equable climate: a new look at the low gradient paradox, J.~Climate, 26, 2117-2136,doi:http://dx.doi.org/10.1175/JCLI-D-11-00547.1, 2013. Sasamori, T.: The radiatice cooling calculation for application to general circulation experiments, J.~Appl.~Meteor., 7, 721-729, 1968. Slingo, A.~and Slingo, J.M.: Response of the National Center for Atmospheric Research community climate model to improvements in the representation of clouds, J.~Geoph.~Res., 96, 341-357, 1991. Sloan, L.C., Walker, J.C.G., and Jr, T.C.M.: Possible role of oceanic heat transport in early Eocene climate, Paleoceanography, 10, 347-356, 1995. Schr\"{o}der, A., Lucarini, V., and Lunkeit, F.: The impact of oceanic heat transport on the atmospheric circulation: a thermodynamic perspective, in prep., 2014. Stephens, G.L.: Radiation profiles in extended water clouds. II: Parameterization schemes, J.~Atmos.~Sci., 34, 2123-2132, 1978. Stephens, G.L., Ackermann, S., and Smith, E.A.: A shortwave parameterization revised to improve cloud absorption, J.~Atmos.~Sci., 41, 687-690, 1984. Stone, P. H.: Constraints on dynamical transports of energy on a spherical planet, Dyn.~Atmos.~Oceans, 2, 123-139, 1978. Ulbrich, U.~and Speth, P.: The global energy cycle of stationary and transient atmospheric waves: Results from ECMWF analyses, Meteorol.~Atmos.~Phys., 45, 125-138, 1991. Winton, M.: On the climate impact on ocean circulation, J.~Climate, 16, 2875-2889, 2003. \clearpage \begin{figure}[t] \centering \includegraphics[width=0.75\textwidth]{fig01.eps} \centering \caption{Oceanic heat transport (in PW) for OHT$_{max}$=0,1,2,3, and 4PW.} \label{fig:OHT} \end{figure} \clearpage \begin{figure}[t] \centering \includegraphics[width=0.75\textwidth]{fig02.eps} \centering \caption{Total heat transport (in PW) diagnosed from energy budget at the top of the atmosphere for OHT$_{max}$=0,1,2,3, and 4PW.} \label{fig:THT} \end{figure} \clearpage \begin{figure}[t] \centering \includegraphics[width=0.75\textwidth]{fig03.eps} \centering \caption{Climatological annual averages for all simulations: Upper: Zonal mean near surface temperature. Lower: Global mean near surface temperatures (T$_M$,in $^{\circ}$C) versus equator-to-pol gradient ($\Delta T$, in $^{\circ}$C).} \label{fig:T2M+TMDT} \end{figure} \clearpage \begin{figure}[t] \centering \includegraphics[width=0.75\textwidth]{fig04.eps} \centering \caption{Climatological southern hemisphere summer (June-August) averages for all simulations: Zonal mean near surface temperatures (in $^{\circ}$C).} \label{fig:T2MJJA} \end{figure} \clearpage \begin{figure}[t] \centering \includegraphics[width=0.75\textwidth]{fig05.eps} \centering \caption{Climatological mean Lorenz energy cycle for June-August: Reservoirs (left, in 10$^5$ J m$^{-2}$) and Conversions (right, in W m$^{-2}$) for yearly data (upper), the winter hemisphere (middle) and the summer hemisphere (lower)for all simulations.} \label{fig:LORENZ} \end{figure} \clearpage \begin{figure}[t] \centering \includegraphics[width=0.75\textwidth]{fig06.eps} \centering \caption{Climatological annual mean mass stream function (in 10$^{10}$ kg s$^{-1}$) for OHT$_{max}$=0,2,3, and 4PW.} \label{fig:MMC} \end{figure} \clearpage \begin{figure}[t] \centering \includegraphics[width=0.75\textwidth]{fig07.eps} \centering \caption{Climatological annual mean mass stream function (northern hemisphere): Strength (in 10$^{10}$ kg s$^{-1}$) and location (in ${^\circ}$N) of Hadley and Ferrel cell for all simulations.} \label{fig:HFC} \end{figure} \clearpage \begin{figure}[t] \centering \includegraphics[width=0.75\textwidth]{fig08.eps} \centering \caption{Climatological annual mean mass stream function (in in 10$^{10}$ kg s$^{-1}$) for OHT$_{max}$=0PW: a) Original (see Figure \ref{fig:MMC}); b) Computed from The Kuo-Eliassen equation (all sources); c) Source from diabatic heating; d) Source from friction; e) Source from eddy heat transport; f) Source from eddy momentum transport.} \label{fig:MMCR} \end{figure} \clearpage \begin{figure}[t] \centering \includegraphics[width=0.75\textwidth]{fig09.eps} \centering \caption{Sources (in 10$^{10}$ kg s$^{-1}$) of the Hadley (upper) and the Ferrel (lower) cell according to the Kuo-Eliassen equation.} \label{fig:MMCF} \end{figure} \end{document}